PCI Architectural Precast Concrete Manual (MNL 122)

PCI_Architectural_Precast_Concrete_Manual

User Manual:

Open the PDF directly: View PDF PDF.
Page Count: 609

DownloadPCI Architectural Precast Concrete Manual (MNL-122)
Open PDF In BrowserView PDF
AR CHITEC TURAL

Precast Concrete

Th ird Edi ti on

Baha’i House of Worship
Wilmette, Illinois;
Architect: Louis Bourgeois.

ARCHITECTURA L
PRECAST CONCRETE
THIRD EDITI ON

209 W. Jackson Boulevard

|

Suite 500

|

Chicago

|

IL 60606-6938

Substantial effort has been made to ensure that all data and information in this manual are
accurate. However, PCI cannot accept responsibility for any errors or oversights in the use of
material or in the preparation of design documens. This publication is intended for use by
personnel competent to evaluate the significance and limitations of its contents and able to
accept responsibility for the application of the material it contains. Special conditions on a
project may require more specific evaluation and practical technical judgment.
While every effort has been made to prepare this publication as best practices standard for
the industry, it is possible that there may be some conflicts between the material herein and
local practices.
MNL-122
Copyright ©2007
By Precast/Prestressed Concrete Institute
First Edition, First Printing, 1973
First Edition, Second Printing, 1974
Second Edition, First Printing, 1989
Third Edition, First Printing, 2007

All rights reserved. This book or any part thereof may not be reproduced in any form without the written permission of the Precast/Prestressed Concrete Institute.
Library of Congress Catalog Card Number 89-062038
ISBN 978-0-937040-78-3

Printed in Canada

Cert no. SW-C0C-1271

The Design Manual for the Architect represents years of intensive work and study within
and outside the Precast/Prestressed Concrete Institute (PCI). The following Committee has
accomplished the task of reflecting and refining these many viewpoints:

P CI A R C H IT E C TU R A L PRECAST CONCRETE COM M ITTEE
SIDNEY FREEDMAN, Editor
GEORGE BATY

ALLAN R. KENNEY

CHARLES L. FISTER

GERVASIO KIM

GARRY FREES

EDWARD S. KNOWLES*

DALE GROFF

JIM LEWIS

DON HALL

CHARLES LOWE

MARVIN F. HARTSFIELD*

PAT O’BRIEN

CATHY HIGGINS

STEVEN P. OTT

TOM HILL

BRUCE D. TAYLOR

LARRY ISENHOUR*

RANDY WILSON

TOM H. KELLEY
*Chair during preparation of current edition

Acknowledgements
The PCI Architectural Precast Concrete Committee wishes to acknowledge the considerable assistance of the following individuals in review and preparation of this Manual: Donald
Benz; Thomas N. Burnham; Richard Fencl; Eve Hinman; Roman Kucharszyk; Jeff LaRue; Ray
A. McCann; Richard Rush; Timothy T. Taylor; Paul Todd; Joseph Trammell; Martha G. Van
Geem – and the many individuals on the PCI Technical Activities Committee, too numerous
to identify, who provided extra effort in their reviews. Special thanks are extended to Jim
Henson and Mark Leader for layout and graphic design.

TABLE OF CONTENTS
n

PREFACE_______________________________________________________________________________________i

CHAPTER ONE – STATE-OF-THE-ART_____________________________________________________________ 1
1.1 MANUAL CONTENT AND CONCEPTS__________________________________________________________ 1
1.2 APPLICATIONS OF ARCHITECTURAL PRECAST CONCRETE_________________________________________ 2
1.3 MISCELLANEOUS USES OF PRECAST CONCRETE________________________________________________ 24
1.4 BENEFITS AND ADVANTAGES OF ARCHITECTURAL PRECAST CONCRETE___________________________ 35
1.5 QUALITY ASSURANCE AND CERTIFICATION PROGRAMS_________________________________________ 38
		 1.5.1 Plant Certification Program_____________________________________________________________ 38
		 1.5.2 Plant Quality Personnel Certification_____________________________________________________ 39
		 1.5.3 Field Certification Program_____________________________________________________________ 39
1.6 DEFINITIONS______________________________________________________________________________ 39
n

CHAPTER TWO – DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS_______________________ 45
2.1 GENERAL COST FACTORS___________________________________________________________________ 45
2.2 DESIGN ECONOMY________________________________________________________________________ 50
		 2.2.1 Repetition___________________________________________________________________________ 50
		 2.2.2 Mold Costs__________________________________________________________________________ 54
		 2.2.3 Other Forming Considerations__________________________________________________________ 56
		 2.2.4 Panel Size and Panelization____________________________________________________________ 58
		 2.2.5 Material and Labor Costs and Uniformity of Appearance____________________________________ 59
		 2.2.6 Design Options______________________________________________________________________ 61
2.3 TOTAL WALL ANALYSIS_____________________________________________________________________ 62
2.4 PRECAST CONCRETE PANELS USED AS CLADDING______________________________________________ 62
		 2.4.1 General_____________________________________________________________________________ 62
		 2.4.2 Solid Wall Panels_____________________________________________________________________ 64
		 2.4.3 Window Wall Panels__________________________________________________________________ 65
		 2.4.4 Spandrel Panels______________________________________________________________________ 66
		 2.4.5 Column Covers and Mullions___________________________________________________________ 68
		 2.4.6 Wall-Supporting Units_________________________________________________________________ 69
2.5 LOADBEARING WALL PANELS OR SPANDRELS__________________________________________________ 73
		 2.5.1 General_____________________________________________________________________________ 73
		 2.5.2 Shapes and Sizes_____________________________________________________________________ 78
		 2.5.3 Design Considerations________________________________________________________________ 81
2.6 PRECAST CONCRETE PANELS USED AS SHEARWALLS_ __________________________________________ 92
2.7 PRECAST CONCRETE AS FORMS FOR CAST-IN-PLACE CONCRETE_________________________________ 94
n

CHAPTER THREE – SURFACE AESTHETICS_ ______________________________________________________ 99
3.1 GENERAL_ _______________________________________________________________________________ 99
3.2 UNIFORMITY AND DEVELOPMENT OF SAMPLES________________________________________________ 99
		 3.2.1 Uniformity_________________________________________________________________________ 100
		 3.2.2 Development of Samples_ ____________________________________________________________ 102
		 3.2.3 Pre-Bid Samples_____________________________________________________________________ 104
		 3.2.4 Production Approval Samples__________________________________________________________ 104
		 3.2.5 Assessment of Samples_______________________________________________________________ 108
		 3.2.6 Assessment of Concrete Mixtures______________________________________________________ 109
3.3 SHAPE, FORM, AND SIZE_ _________________________________________________________________ 111
		 3.3.1 Open or Closed Units_ _______________________________________________________________ 111
		 3.3.2 Drafts_ ____________________________________________________________________________ 112
		 3.3.3 Reveals and Demarcation Features_ ____________________________________________________ 113
		 3.3.4 Sculpturing_________________________________________________________________________ 120
		 3.3.5 Bullnoses, Arrises and Radiused Precast Concrete_________________________________________ 125
		 3.3.6 Cornices and Eyebrows_______________________________________________________________ 131
		 3.3.7 Edges, Corners and Returns___________________________________________________________ 131
		 3.3.8 Returns in Relation to Finishes_________________________________________________________ 138
n

ARCHITECTURAL PRECAST CONCRETE

		 3.3.9 Two-Stage or Sequential Precasting_____________________________________________________ 139
		 3.3.10 Overall Panel Size__________________________________________________________________ 140
3.4 COLORS AND TEXTURES___________________________________________________________________ 144
		 3.4.1 Colors_____________________________________________________________________________ 144
		 3.4.2 Textures____________________________________________________________________________ 150
3.5 FINISHES________________________________________________________________________________ 153
		 3.5.1 General____________________________________________________________________________ 153
		 3.5.2 Smooth-As-Cast_____________________________________________________________________ 153
		 3.5.3 Exposed Aggregate by Chemical Retarders and Water Washing_____________________________ 156
		 3.5.4 Form Liners and Lettering_____________________________________________________________ 160
		 3.5.5 Sand or Abrasive Blasting_____________________________________________________________ 169
		 3.5.6 Acid Etching________________________________________________________________________ 174
		 3.5.7 Multiple Mixtures and Textures within a Single Unit_______________________________________ 177
		 3.5.8 Tooling or Bushhammering____________________________________________________________ 182
		 3.5.9 Sand Embedment_ __________________________________________________________________ 186
		 3.5.10 Clay Product-Faced Precast Concrete_ _________________________________________________ 187
					 3.5.10.1 General___________________________________________________________________ 187
					 3.5.10.2 General considerations______________________________________________________ 188
					 3.5.10.3 Clay product properties______________________________________________________ 190
					 3.5.10.4 Clay product selection_______________________________________________________ 190
					 3.5.10.5 Design considerations_______________________________________________________ 200
					 3.5.10.6 Production and construction considerations_____________________________________ 203
					 3.5.10.7 Applications of clay products after casting of panel_ _____________________________ 204
		 3.5.11 Honed or Polished__________________________________________________________________ 206
		 3.5.12 Stone Veneer-Faced Precast Concrete__________________________________________________ 211
					 3.5.12.1 General considerations______________________________________________________ 211
					 3.5.12.2 Stone properties____________________________________________________________ 212
					 3.5.12.3 Stone sizes________________________________________________________________ 213
					 3.5.12.4 Design considerations_______________________________________________________ 214
					 3.5.12.5 Anchorage of stone facing___________________________________________________ 214
					 3.5.12.6 Panel watertightness________________________________________________________ 217
					 3.5.12.7 Veneer jointing_____________________________________________________________ 217
					 3.5.12.8 Repair____________________________________________________________________ 218
					 3.5.12.9 Applications_______________________________________________________________ 218
		 3.5.13 Applied Coatings___________________________________________________________________ 229
		 3.5.14 Architectural Trim Units_____________________________________________________________ 231
		 3.5.15 Matching of Precast and Cast-in-Place Concrete_________________________________________ 238
		 3.5.16 Finishing of Interior Panel Faces_______________________________________________________ 239
		 3.5.17 Acceptability of Appearance_ ________________________________________________________ 240
		 3.5.18 Repair and Patching_ _______________________________________________________________ 241
3.6 WEATHERING____________________________________________________________________________ 242
		 3.6.1 General____________________________________________________________________________ 242
		 3.6.2 Surface Finish_______________________________________________________________________ 248
		 3.6.3 Deposits From an Adjacent Surface or Material___________________________________________ 250
		 3.6.4 Efflorescence on Precast Concrete______________________________________________________ 250
				3.6.4.1 What is efflorescence_ ________________________________________________________ 251
				3.6.4.2 What causes efflorescence_____________________________________________________ 252
				3.6.4.3 Minimizing efflorescence_ _____________________________________________________ 253
		 3.6.5 Design of Concrete for Weathering_____________________________________________________ 255
		 3.6.6 Surface Coatings and Sealers__________________________________________________________ 256
		 3.6.7 Maintenance and Cleaning___________________________________________________________ 259
CHAPTER FOUR – DESIGN_ ___________________________________________________________________ 263
4.1 DESIGN AND CONSTRUCTION RESPONSIBILITY________________________________________________ 263
		 4.1.1 General____________________________________________________________________________ 263
n

ARCHITECTURAL PRECAST CONCRETE

TABLE OF CONTENTS

		 4.1.2 Responsibilities of the Architect________________________________________________________ 263
		 4.1.3 Responsibilities of the Engineer of Record_ ______________________________________________ 266
		 4.1.4 Responsibilities of the General Contractor / Construction Manager__________________________ 267
			
4.1.4.1 Bid process_ _________________________________________________________________ 268
		 4.1.5 Responsibilities of the Precaster________________________________________________________ 269
		 4.1.6 Responsibilities of the Erector__________________________________________________________ 270
4.2 STRUCTURAL DESIGN_____________________________________________________________________ 271
		 4.2.1 General Considerations_______________________________________________________________ 271
			
4.2.1.1 Design objectives_____________________________________________________________ 271
			
4.2.1.2 Design criteria________________________________________________________________ 272
			
4.2.1.3 Checklist____________________________________________________________________ 273
		 4.2.2 Determination of Loads______________________________________________________________ 274
		 4.2.3 Volume Changes____________________________________________________________________ 275
			
4.2.3.1 Temperature effects___________________________________________________________ 275
			
4.2.3.2 Shrinkage_ __________________________________________________________________ 275
			
4.2.3.3 Creep_______________________________________________________________________ 276
		 4.2.4 Design Considerations for Non-Loadbearing Wall Panels___________________________________ 276
			
4.2.4.1 Deformations_ _______________________________________________________________ 277
			
4.2.4.2 Column covers and mullions____________________________________________________ 280
		 4.2.5 Design Considerations for Loadbearing Wall Panels_______________________________________ 281
		 4.2.6 Design Considerations for Non-Loadbearing Spandrels_ ___________________________________ 282
		 4.2.7 Design Considerations for Loadbearing Spandrels_________________________________________ 285
		 4.2.8 Design Considerations for Stacking Non-Loadbearing Panels_______________________________ 286
		 4.2.9 Dimensioning of Precast Concrete Units_________________________________________________ 287
		 4.2.10 Handling and Erection Considerations_ ________________________________________________ 290
			
4.2.10.1 Wall panels_________________________________________________________________ 295
			
4.2.10.2 Columns___________________________________________________________________ 296
			
4.2.10.3 Spandrels___________________________________________________________________ 297
			
4.2.10.4 Column covers and mullions___________________________________________________ 297
			
4.2.10.5 Soffits_ ____________________________________________________________________ 297
			
4.2.10.6 Protection during erection_____________________________________________________ 297
4.3 CONTRACT DOCUMENTS__________________________________________________________________ 298
4.4 REINFORCEMENT_________________________________________________________________________ 300
		 4.4.1 General____________________________________________________________________________ 300
		 4.4.2 Welded Wire Reinforcement_ _________________________________________________________ 301
		 4.4.3 Reinforcing Bars_____________________________________________________________________ 302
		 4.4.4 Prestressing Steel____________________________________________________________________ 302
		 4.4.5 Shadow Lines–Reflection of Reinforcing Steel____________________________________________ 304
		 4.4.6 Tack Welding_______________________________________________________________________ 305
		 4.4.7 Corrosion Resistance of Reinforcement__________________________________________________ 305
			
4.4.7.1 Chlorides____________________________________________________________________ 306
			
4.4.7.2 Concrete cover_______________________________________________________________ 306
			
4.4.7.3 Permeability_________________________________________________________________ 307
			
4.4.7.4 Carbonation_________________________________________________________________ 307
			
4.4.7.5 Crack widths_________________________________________________________________ 309
			
4.4.7.6 Corrosion protection__________________________________________________________ 310
4.5 CONNECTIONS___________________________________________________________________________ 312
		 4.5.1 General____________________________________________________________________________ 312
			
4.5.1.1 Design Responsibilities_________________________________________________________ 313
		 4.5.2 Design Considerations_______________________________________________________________ 313
			
4.5.2.1 Panel configuration_ __________________________________________________________ 315
			
4.5.2.2 Panel–connection–structure interaction___________________________________________ 316
			
4.5.2.3 Tolerances and product interfacing_ _____________________________________________ 320
			
4.5.2.4 Other detailing information____________________________________________________ 321

ARCHITECTURAL PRECAST CONCRETE

		 4.5.3 Handling and Erection Considerations_ _________________________________________________ 322
		 4.5.4 Handling and Lifting Devices_ _________________________________________________________ 324
		 4.5.5 Manufacturing Considerations_________________________________________________________ 325
		 4.5.6 Connection Hardware and Materials____________________________________________________ 325
		 4.5.7 Corrosion Protection of Connections_ __________________________________________________ 326
		 4.5.8 Fasteners in Connections_ ____________________________________________________________ 327
		 4.5.9 Supply of Hardware for Connections_ __________________________________________________ 330
		 4.5.10 Connection Details_ ________________________________________________________________ 331
4.6 TOLERANCES____________________________________________________________________________ 347
		 4.6.1 General____________________________________________________________________________ 347
		 4.6.2 Product Tolerances___________________________________________________________________ 347
		 4.6.3 Erection Tolerances_ _________________________________________________________________ 350
		 4.6.4 Interfacing Tolerances________________________________________________________________ 363
4.7 JOINTS _ ________________________________________________________________________________ 364
		 4.7.1 General____________________________________________________________________________ 364
		 4.7.2 Types of Joints_ _____________________________________________________________________ 365
		 4.7.3 Expansion Joints_____________________________________________________________________ 366
		 4.7.4 Number of Joints____________________________________________________________________ 368
		 4.7.5 Location of Joints____________________________________________________________________ 368
		 4.7.6 Width and Depth of Joints____________________________________________________________ 369
		 4.7.7 Sealant Materials and Installation_ _____________________________________________________ 371
		 4.7.8 Architectural Treatment_______________________________________________________________ 375
		 4.7.9 Fire-Protective Treatment_ ____________________________________________________________ 376
		 4.7.10 Joints in Special Locations___________________________________________________________ 376
CHAPTER FIVE – OTHER ARCHITECTURAL DESIGN CONSIDERATIONS_____________________________ 379
5.1 GENERAL_ ______________________________________________________________________________ 379
5.2 WINDOWS AND GLAZING_________________________________________________________________ 379
		 5.2.1 Design Considerations_______________________________________________________________ 379
		 5.2.2 Window Installation_ ________________________________________________________________ 384
		 5.2.3 Other Attached or Incorporated Materials_______________________________________________ 386
		 5.2.4 Glass Staining or Etching_ ____________________________________________________________ 387
5.3 ENERGY CONSERVATION AND CONDENSATION CONTROL______________________________________ 390
		 5.3.1 Glossary_ __________________________________________________________________________ 390
		 5.3.2 Energy Conservation_________________________________________________________________ 391
		 5.3.3 Thermal Resistance (R-Value)__________________________________________________________ 402
		 5.3.4 Heat Capacity_______________________________________________________________________ 410
		 5.3.5 Thermal Mass_______________________________________________________________________ 412
		 5.3.6 Condensation Control________________________________________________________________ 417
			
5.3.6.1 Climates____________________________________________________________________ 418
			
5.3.6.2 Sources of moisture___________________________________________________________ 418
			
5.3.6.3 Condensation on surfaces______________________________________________________ 420
			
5.3.6.4 Condensation within walls and use of vapor retarders_ _____________________________ 423
			
5.3.6.5 Air infiltration, exfiltration, and air barriers________________________________________ 444
			
5.3.6.6 Considerations at windows_____________________________________________________ 447
		 5.3.7 Application of Insulation______________________________________________________________ 447
		 5.3.8 Precast Concrete Sandwich Panels______________________________________________________ 450
5.4 SUSTAINABILITY__________________________________________________________________________ 459
		 5.4.1 Glossary_ __________________________________________________________________________ 459
		 5.4.2 Sustainability Concepts_______________________________________________________________ 460
			
5.4.2.1 Triple bottom line_____________________________________________________________ 461
			
5.4.2.2 Cost of building green_________________________________________________________ 461
			
5.4.2.3 Holistic/integrated design_ _____________________________________________________ 462
			
5.4.2.4 3 R’s–reduce, reuse, recycle_____________________________________________________ 463
n

ARCHITECTURAL PRECAST CONCRETE

TABLE OF CONTENTS

		 5.4.3 Life Cycle_ _________________________________________________________________________ 463
			
5.4.3.1 Life cycle cost and service life___________________________________________________ 463
			
5.4.3.2 Environmental life cycle inventory and life cycle assessment__________________________ 464
			
5.4.3.3 Concrete and concrete products LCI_ ____________________________________________ 465
			
5.4.3.4 Life cycle impact assessment____________________________________________________ 467
		 5.4.4 Green Building Rating Systems_________________________________________________________ 469
			
5.4.4.1 LEED________________________________________________________________________ 469
			
5.4.4.2 Energy Star__________________________________________________________________ 470
			
5.4.4.3 Green Globes________________________________________________________________ 472
		 5.4.5 Durability_ _________________________________________________________________________ 473
			
5.4.5.1 Corrosion resistance___________________________________________________________ 473
			
5.4.5.2 Inedible_____________________________________________________________________ 473
		 5.4.6 Resistant to Natural Disasters__________________________________________________________ 473
			
5.4.6.1 Fire Resistance________________________________________________________________ 473
			
5.4.6.2 Tornado, hurricane, and wind resistance__________________________________________ 474
			
5.4.6.3 Flood resistance_ _____________________________________________________________ 474
			
5.4.6.4 Earthquake resistance_ ________________________________________________________ 474
		 5.4.7 Weather Resistance__________________________________________________________________ 474
			
5.4.7.1 High humidity and wind-driven rain______________________________________________ 474
			
5.4.7.2 Ultraviolet resistance_ _________________________________________________________ 475
		 5.4.8 Mitigating the Urban Heat Island Effect_________________________________________________ 475
			
5.4.8.1 Warmer surface temperatures__________________________________________________ 475
			
5.4.8.2 Smog_______________________________________________________________________ 475
			
5.4.8.3 Albedo (solar reflectance)______________________________________________________ 475
			
5.4.8.4 Emittance_ __________________________________________________________________ 477
			
5.4.8.5 Moisture_ ___________________________________________________________________ 477
			
5.4.8.6 Mitigation approaches_________________________________________________________ 477
			
5.4.8.7 Thermal mass and nocturnal effects_____________________________________________ 477
		 5.4.9 Environmental Protection_____________________________________________________________ 477
			
5.4.9.1 Radiation and toxicity_________________________________________________________ 477
			
5.4.9.2 Resistance to noise____________________________________________________________ 478
			
5.4.9.3 Security and impact resistance__________________________________________________ 478
		 5.4.10 Precast Concrete Production_ ________________________________________________________ 478
					 5.4.10.1 Constituent materials_ ______________________________________________________ 479
		 5.4.11 Energy Use in Buildings______________________________________________________________ 482
			
5.4.11.1 Energy codes_______________________________________________________________ 482
			
5.4.11.2 Lighting____________________________________________________________________ 484
			
5.4.11.3 Air infiltration_______________________________________________________________ 484
			
5.4.11.4 Advanced energy guidelines___________________________________________________ 484
		 5.4.12 Indoor Environmental Quality_________________________________________________________ 485
		 5.4.13 Demolition________________________________________________________________________ 486
		 5.4.14 Innovation_ _______________________________________________________________________ 486
		 5.4.15 Conclusion________________________________________________________________________ 487
5.5 ACOUSTICAL PROPERTIES_ ________________________________________________________________ 487
		 5.5.1 Glossary_ __________________________________________________________________________ 487
		 5.5.2 General____________________________________________________________________________ 488
		 5.5.3 Sound Levels_______________________________________________________________________ 488
		 5.5.4 Sound Transmission Loss______________________________________________________________ 489
		 5.5.5 Absorption of Sound_________________________________________________________________ 489
		 5.5.6 Acceptable Noise Criteria_____________________________________________________________ 491
		 5.5.7 Composite Wall Considerations________________________________________________________ 493
		 5.5.8 Leaks and Flanking_ _________________________________________________________________ 496
5.6 DESIGN CONSIDERATIONS FOR BLAST RESISTANCE____________________________________________ 497
		 5.6.1 General____________________________________________________________________________ 497

ARCHITECTURAL PRECAST CONCRETE

		 5.6.2 Blast Basics_________________________________________________________________________ 498
		 5.6.3 Blast Analyses Standards______________________________________________________________ 499
		 5.6.4 Determination of Blast Loading________________________________________________________ 500
		 5.6.5 Blast Effects Predictions_______________________________________________________________ 501
		 5.6.6 Standoff Distance_ __________________________________________________________________ 504
		 5.6.7 Design Concepts____________________________________________________________________ 506
		 5.6.8 Façade Considerations_ ______________________________________________________________ 509
		 5.6.9 Designing Precast Concrete Panels_ ____________________________________________________ 509
		 5.6.10 Examples of Projects Designed for Blast________________________________________________ 513
		 5.6.11 Connection Concepts and Details_____________________________________________________ 517
		 5.6.12 Glazing___________________________________________________________________________ 520
		 5.6.13 Initial Costs________________________________________________________________________ 523
		 5.6.14 References________________________________________________________________________ 524
5.7 FIRE RESISTANCE_________________________________________________________________________ 525
		 5.7.1 General____________________________________________________________________________ 525
		 5.7.2 Fire Endurance of Walls_______________________________________________________________ 527
		 5.7.3 Detailing of Fire Barriers______________________________________________________________ 532
		 5.7.4 Columns and Column Covers_________________________________________________________ 533
		 5.7.5 Protection of Reinforcing Steel_________________________________________________________ 535
		 5.7.6 Protection of Connections_ ___________________________________________________________ 535
5.8 ROOFING_ ______________________________________________________________________________ 536
		 5.8.1 General____________________________________________________________________________ 536
		 5.8.2 Flashing____________________________________________________________________________ 536
		 5.8.3 Parapet Details______________________________________________________________________ 539
		 5.8.4 Scuppers___________________________________________________________________________ 542
CHAPTER SIX – GUIDE SPECIFICATIONS________________________________________________________ 545
6.1 GENERAL_ ______________________________________________________________________________ 545
6.2 DRAWINGS AND SPECIFICATIONS___________________________________________________________ 545
		 6.2.1 Drawings__________________________________________________________________________ 545
		 6.2.2 Specifications_______________________________________________________________________ 545
		 6.2.3 Coordination_ ______________________________________________________________________ 545
		 6.2.4 Guide Specification Development______________________________________________________ 546
6.3 TYPES OF SPECIFICATIONS_________________________________________________________________ 546
6.4 GUIDE SPECIFICATION_____________________________________________________________________ 547
PART 1 – GENERAL_ ______________________________________________________________________ 547
PART 2 – PRODUCTS______________________________________________________________________ 556
PART 3 – EXECUTION_ ____________________________________________________________________ 575
n

n

INDEX BY SUBJECTS_ ________________________________________________________________________ 581

ARCHITECTURAL PRECAST CONCRETE

PREFACE

A

rchitectural precast concrete is a child of the 20th
century and modern technology, but it can trace
its lineage back to ancient history. As such, it is a
building material almost without precedent. Concrete
in its cruder forms was used by the Romans in the
construction of their aqueducts. Europe refined the
time-tested formula in the 19th century, developing
reinforced concrete that combined the compressive
properties of concrete and the tensile strength of steel.
Continuing technological growth and industrialization
created a genuine need for new techniques and materials that could be used in prefabricated construction.
Architectural precast concrete was developed to fulfill
this need.

The first documented modern use of precast concrete was
in the cathedral Notre Dame du Raincy in Raincy, France, by
Auguste Perret in 1922. It was used as screen walls and
infill in an otherwise in-situ concrete solution. In 1932 work
began on producing the white concrete exposed aggregate ornamental elements for the Baha’i House of Worship
(frontispiece and Fig. 1.2.1). The Depression years followed
soon after, and then World War II. Following the end of
the world conflict, when labor and material costs began to
increase, the use of architectural precast concrete began
to flourish. The development and introduction of improved
transportation equipment and large tower cranes on major projects provided a ready means of hauling and lifting
large precast concrete panels into place. By the mid-1960s,
architectural precast concrete as cladding and loadbearing
elements had gained widespread acceptance by architects
and owners.
Improvements in fabricating processes allow architectural
precast concrete to be produced in almost any color, form,
or texture, making it an eminently practical and aesthetically
pleasing building material. The term architectural precast
concrete encompasses all precast concrete units employed
as elements of architectural design whether defined to stand
alone as an architectural statement or to complement other
building materials. Concrete’s moldability offers the freedom to sculpt the structure’s facade in imaginative ways.
It is difficult to imagine an architectural style that cannot
be expressed with this material. Precast concrete is not only
compatible with all structural systems, it can be designed to
harmonize with, and complement, all other materials.
Throughout the formative years, the architect, the engineer, and the builder, as well as the precaster, lacked any
definitive reference volume that defined and illustrated
this interesting material. This lacking was both world- and
language-wide.
The Precast/Prestressed Concrete Institute (PCI), a nonprofit corporation founded in 1954 to advance the design,
manufacture, and use of precast and prestressed concrete,

i

|

ARCHITECTURAL PRECAST CONCRETE

had long recognized the need for a manual to provide
guidelines and recommendations pertinent to the design,
detailing, and specification of architectural precast concrete.
In 1973, PCI published the first edition of Architectural
Precast Concrete and for the first time there was a comprehensive design manual on the subject of architectural
precast concrete. Compiled, edited, and published by PCI,
this manual presented a single authoritative reference for
the architectural decision-maker.
New developments in materials, manufacturing, and erection procedures have expanded the role of architectural
precast concrete in the construction industry since the first
manual was written. In keeping with its policy of being a
leader in the concrete industry, PCI is publishing this third
edition of Architectural Precast Concrete in order to make
state-of-the-art technology available to the architects and
engineers who design and build with this versatile material.
The third edition of the manual is a major revision incorporating much of this new technology. The sections dealing with color, texture, and finishes; weathering; tolerances;
connections; and thermal properties have been extensively
revised. Information on sustainability and design for blast
has been added. Detailed guide specifications have been
modified to meet today’s construction needs. In addition, the photographs used to illustrate pertinent points
throughout the manual have been selected to represent
the potential design opportunities for architectural precast
concrete.
Numerous manufacturing and erection techniques are
included in the text to provide a better understanding of
design concepts and elements requiring design decisions.
Design, contract drawings, and specifications are all vitally
important, and should be combined with an assessment of
the capability and experience of the precasters who bid on
the project.
The guidelines and recommendations presented show
current practices in the industry. These practices should not,
however, act in any way as barriers to either architectural creativity or to potential innovations on the part of the
precaster.
The practices described in this manual may be used as a
basis by both architect and precaster in the development of
exciting new concepts using advances in technology. This
initiative will undoubtedly lead to deviations from some of
the stated recommendations in the text.
The editor of the manual worked closely with the PCI
Architectural Precast Concrete Committee and with the
professional members of the Institute. Technical accuracy
has been reviewed by architects, engineers, precast concrete producers, material and equipment suppliers, and af-

filiated industry organizations. This unique combination of
various disciplines and viewpoints provides an interaction
that ensures knowledge of all aspects in design, engineering, production, and erection of architectural precast concrete. Since conditions affecting the use of this material are
beyond the control of the Institute, the suggestions and
recommendations presented are provided without guarantee or responsibility on the part of PCI. It is assumed that
each project and architect is unique, and requires different
solutions for different problems. For this reason, all examples shown must be considered as suggestions rather than
definitive solutions.
Architectural precast concrete combines maximum freedom of architectural expression with the economy of mass
production of repetitive precast concrete elements. For this
concept to function most effectively, it is strongly recommended that the architect seek counsel from local PCI and
Canadian Prestressed Concrete Institute (CPCI) architectural precast concrete producers in the early design stages
and throughout further development of the contract documents (see PCI website www.pci.org for local producers
and resources). Many consulting engineering firms specializing in the development and design of precast concrete
are also available to the project architect.
With precaster/consultant assistance, proper aesthetic,
functional, structural, mechanical features and objectives
may be rendered with economical detailing. Their assistance
may more accurately reflect local material characteristics,
manufacturing and erection efficiencies, cost factors, quality control standards, and local trade practices. A continuing
dialogue between designer and precaster will ensure optimum product quality at a minimum installed construction
cost.
This manual is arranged in a sequence that corresponds
to the steps that an architectural/engineering firm might
employ when evaluating, selecting, and incorporating materials into a construction project. Other publications of
interest to the design team include PCI Design Handbook
– Precast and Prestressed Concrete (MNL-120); PCI Manual
for Quality Control for Plants and Production of Architectural
Precast Concrete Products (MNL-117); PCI Erector’s Manual
– Standards and Guidelines for the Erection of Precast
Concrete Products (MNL-127); and PCI Erection Safety for
Precast and Prestressed Concrete (MNL-132).
Chapter 1 provides a general background concerning
the State-of-the-Art of architectural precast concrete and
covers the applications and benefits of architectural precast
concrete along with definitions.
Chapter 2 considers Design Concepts Related to
Usage and Economics for the initial evaluation and selection of architectural precast concrete for a project. The

architect would primarily use this information during conceptual wall analysis. Repetition and the master mold concept portions of Chapter 2 would be of most interest to the
architect’s staff concerned with production and detailing.
Chapter 3 contains Surface Aesthetics design considerations and is concerned with the critical decisions the
designer must make among the many available options as
to color, shape, and texture. This chapter covers everything
from initial samples and concrete mixture design to acceptability of appearance and weathering. Weathering should
be reviewed during conceptual wall analysis.
Chapter 4 presents Design. This chapter covers design
responsibilities, impact of structural framing considerations,
contract documents, reinforcement, connections, tolerances, and joints. The job captain, draftsman, and detailer
members of the architect’s staff must carefully consider
these factors. They should also be familiar with the design
considerations included in Chapters 2 and 3 if a sound,
economical finished product is to result.
Chapter 5 reviews Other Architectural Design Considerations and covers interfacing with other materials
including windows, energy conservation and condensation
control, sustainability, acoustical properties, blast considerations, fire resistance, and roofing. Each requires careful
consideration in developing the design criteria and working
details for the architectural precast concrete systems and
assemblies adjacent to the precast concrete.
Chapter 6 Guide Specifications is intended as an aid
to specification writers. The information contained in this
chapter should be evaluated in close coordination with the
project designer and detailer to avoid creating unnecessary
pitfalls in the project by providing the best possible contract
documents. Specifications should be neither open to interpretation nor unnecessarily restrictive.
An Index is provided at the end of the manual for easy
reference.
A design manual by its very concept can only illustrate
what has been accomplished, not what can be. Any attempt to categorize and define architectural precast concrete with its myriad expressions and possibilities is not fully
possible. Precast concrete is a versatile material that offers
the designer the opportunity to be innovative and obtain
desired design objectives that cannot be accomplished with
other materials. This manual will help architects define their
own potential and will provide a basis for reaching it, not
by giving design alternatives, but by pointing out state-ofthe-art options in using architectural precast concrete. The
significance of precast concrete as a building material lies
not in its ability to do new things, but in its inherent quality
of being flexible enough to make a design concept become
a reality.

ARCHITECTURAL PRECAST CONCRETE

|

ii

1

Santa Monica
Public Library
Santa Monica, California;
Architect: Moore Ruble
Yudell Architects &
Planners; Photo:
John Edward Linden
Photography.

CHAPTER ONE
STATE-OF-THE- A RT
1.1 M
 ANUAL CONTENT AND CONCEPTS

considerations dictated by:

In order for the reader to derive maximum benefit
from this manual, the concepts that guided its preparation are briefly outlined in this section.

1. Material characteristics;

Concrete is material that has been used for centuries
but its present qualities and usages characterize it as
modern and versatile. The color and texture possibilities along with the infinite variety of forms and shapes
possible with architectural precast concrete, as a result of its plasticity, allows the designer to customize
elements and overcome the monotony of many proprietary systems, the confining regularity of masonry
units, and the design limitations of materials subject
to the shortcomings of dies, brakes, and rollers. With
precast concrete, the designer is in complete control
of the ultimate form of the façade and is free of any
compromise that might be traceable to a manufacturing process. Also, precast concrete has excellent inherent thermal, acoustical, fire resistive and blast resistant
characteristics.

4. Colors, finishes, and textures; and

Research and quality control of concrete components
has led to a better understanding of the unique potentials for precast concrete. Improvements in proportioning, mixing, placing, finishing, and curing techniques
have advanced concrete qualities. These improvements have allowed design strength to be substantially
increased in the last few decades. Durability, appearance, and other important aspects, such as transportation and erection, have kept pace with these
developments.
By continuing to provide the designer with design
freedom, the precast concrete industry has experienced
steady growth since the 1960s, particularly in the everwidening range of precast concrete applications. The
widespread availability of architectural precast concrete, the nearly universal geographic distribution of
the necessary raw materials, and the high construction
efficiency of prefabricated components all add to the
appeal of architectural precast concrete construction.
Precast concrete design engineers and precasters have
a high level of craftsmanship and ingenuity along with
a thorough knowledge of the material and its potential in converting the designer’s vision into a finished
structure. Related to this knowledge must be an understanding by the designer of the important design

2. Wall analysis — interrelationship with other materials;
3. Weathering effects;
5. Manufacturing and erection efficiencies including
cost factors, quality-control standards, and local
trade practices.
These design considerations are described in considerable detail throughout this manual. It is important
that the designer carefully evaluate the applicable design considerations when choosing the shape, color,
and texture that will be emphasized on a project.
Design considerations based upon manufacturing and
erection factors are complex and varied. Because of
the limitless expressions of precast concrete, such design considerations can rarely be stated in unqualified
terms. It is important for the designer to know which
considerations are valid and to assess their influence
on a specific application. The design considerations for
architectural precast concrete are not any more numerous or difficult than the ones associated with other
materials. Optimum utilization of precast concrete will
result from a thorough understanding of the applicable design considerations. Consider, for example, the
interplay between the configuration of a precast concrete unit and its structural capacity, and between its
shape and available finishes. Add to this the available
options in reinforcement, including the possibility of
prestressing the unit to offset tensile stresses caused by
handling or service conditions, and the importance of
understanding these design considerations becomes
clear.
The architect, with or without prior experience with
precast concrete, will benefit from a detailed study of
the entire manual in order to obtain an understanding
of interrelated design considerations. Subsequent to
such a study, the manual will be a valuable reference
for the future design of architectural precast concrete.
Structural precast concrete products comprise an important segment of the precast concrete market. These
units are normally produced in standard shapes such
as double or single tees, and channel, solid, or hollow slabs with many different finishes that are often

ARCHITECTURAL PRECAST CONCRETE

|

1

1

STATE-OF-THE-ART

1.1 Manual Content And Concepts / 1.2 Applications of Architectural Precast Concrete

machine applied. Although many of the design considerations discussed in this manual apply to these products, additional design considerations may also govern.
These design considerations are thoroughly covered in
the companion PCI Design Handbook—Precast and
Prestressed Concrete. Similarly, information concerning regional availability of products and product variation between individual plants is readily available from
local producers. Thus, structural precast concrete products are not specifically discussed in this manual. This
manual concentrates on the architectural precast concrete applications that are custom designed in shapes
and finishes for each individual project.
The manual contains some recommendations with respect to job requirements and contract conditions relating
to precast concrete. These procedural recommendations

will help to minimize complications and facilitate communications during the bidding and construction stages.
The design considerations, together with the procedural recommendations, should be geared to local
practices. The importance of coordinating the development of architectural precast concrete projects with local precasters cannot be overemphasized. The ultimate
aim of this manual is to encourage this communication
and forming of relationships.

1.2 APPLICATIONS OF ARCHITECTURAL
PRECAST CONCRETE
Architectural precast concrete has been used for many
decades in North America. Its full potential in terms of

Fig. 1.2.1 Interior of Baha’i House of Worship, Wilmette, Illinois; Architect: Louis Bourgeois.

2

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

Fig. 1.2.2 Jefferson County Government Center, Golden, Colorado; Architect: C.W. Fentress and J.H. Bradburn Associates.

economy, versatility, appearance, structural strength,
quality, and performance continues to expand as witnessed by the new projects changing the skylines of
many cities throughout the North American continent.
Even today, buildings with precast concrete cladding
dating back to the 1920s and 1930s attest to the fine
craftsmanship of those periods and the permanence of
the material. A classic example is the Baha’i House of
Worship, Wilmette, Illinois (frontispiece and Fig. 1.2.1),
designed by Canadian architect Louis Bourgeois. This
structure, started in 1920 with final completion in
1953, is one of the most beautiful and delicately detailed structures ever constructed in the United States.
It is a nine-sided structure built with white architectural precast concrete panels with exposed quartz
aggregate over a steel superstructure. Each side has
the form of a circular arc, with a large doorway in the
center. Pylons 45 ft (13.7 m) in height stand at the corners of the first story. Above the gallery, the clerestory
and the dome are also nine-sided but with the ribs rising from midway of the sides of the first story. On the
dome is inscribed the orbits of the stars and planets in
a pattern of ovals, circles, and flowing curves. Symbols
representing life, tendrils, flowers, leaves, and fruit are
woven into the design. The interior (Fig. 1.2.1) is a lofty
cylindrical room topped with a hemispherical dome of
75 ft (22.9 m) interior diameter and extending to a
height of 135 ft (41.1 m) in the center. Wire glass is
supported by a steel framework concealed within the
intricately patterned precast concrete exterior and interior surface. These act as perforated screens through
which light passes.

Today, architectural precast concrete demands equal
craftsmanship in the design and tooling aspects of the
manufacturing process. Production has progressed
from reliance on individual craftsmanship to a well controlled and coordinated production line method with
corresponding economic and physical improvements.
These state-of-the-art manufacturing techniques do
not sacrifice the plastic qualities of concrete, nor do
they limit the freedom of three-dimensional design. In
knowledgeable, sympathetic hands, these techniques
can be adapted to fit specific performance and aesthetic requirements of the contemporary designer.
Treatment of texture and color can be rich, extracting the best qualities of the raw materials. Coarse aggregates are selected for their size, shape, and color.
Fine aggregates and cement are selected according to
the desired texture and color of the finished element.
The mixing of aggregates and cement is similar to the
artist’s mixing of colors on a palette.
The photographs in this chapter afford a glimpse of
the variety of expressions possible with precast concrete; bold yet simple to emphasize strength and durability; intricate as well as delicate to mirror elegance;
and three-dimensional sculpturing to display individuality. Other photographs or drawings throughout the
manual illustrate these varieties of expressions along
with specific concepts and details.
Complex molds with 30 different radii were constructed to produce more than 4000 precast concrete
components that form the shell for the structure in Fig.
1.2.2. Two colors of precast concrete were used—a tan

ARCHITECTURAL PRECAST CONCRETE

|

3

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

and an earth-tone burnt orange. Both were selected
for the compatibility in color and hue with the surrounding landscape and terrain.
The building in Fig. 1.2.3 features a unique profile, with
a two-sided curved penthouse descending to a top floor
curved only on one side, which is juxtaposed with the rotated square tower footprint below. The design creates
the impression that a curving, modern structure is blending with a traditional rectangular one. Both architectural
precast concrete panels and granite are used to clad the
30-story office building. The first three floors are clad with
flame-finish granite anchored to precast concrete backing
panels with horseshoe-shaped stainless steel pin anchors.
The upper floors feature precast concrete panels that have
been finely detailed and textured with sandblasting and
the application of a retarder finish. This treatment provides
the panels with a flamed, stone-like appearance that provides a seamless and virtually indiscernible shift from granite to concrete.

Fig. 1.2.4(a) Police
Administration Building
Philadelphia, Pennsylvania;
Architect: Cubellis GBQC formerly Geddes Brecher Qualls
Cunningham, Architects; Photos:
Portland Cement Association.
Fig. 1.2.4(b)
Loadbearing wall construction.

Fig. 1.2.3 IJL Financial Center, Charlotte, North Carolina; Architect:
Smallwood, Reynolds, Stewart, Stewart & Associates, Inc.

4

|

ARCHITECTURAL PRECAST CONCRETE

The Police Administration
Building in Philadelphia made
history in 1961 as one of the
first major buildings to use
the inherent structural characteristics of architectural
precast concrete (Fig. 1.2.4[a] and [b]). The 5-ft-wide (1.5
m), 35-ft-high (10.7 m), three-story exterior panels carry
two upper floors and the roof. The building is unusual in
its plan configuration, consisting of two circles connected
by a curving central section, demonstrating the adaptability of concrete to unusual floor plans. This structure was

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.5 U.S. Department of Housing and Urban Development
Headquarters, Washington, D.C.; Architect: Marcel Breuer
and Herbert Beckhard Design Architects; Nolen-Swinburne &
Associates, Associate Architects; Photo: ©Images are courtesy
of the Marcel Breuer papers, 1920-1986, in the Archives of
American Art, Smithsonian Institution.

an early model for the blending of multiple systems into
one building. Precast concrete and post-tensioning were
the relatively new techniques that were then successfully
combined.
The headquarters building for the Department of
Housing and Urban Development in Washington, D.C.,
completed in 1968, has loadbearing panels that house
air-conditioning units below sloping sills and form vertical
chases for mechanical services (Fig. 1.2.5). The windows
are recessed in the deeply sculptured panels for solar control. The panels lend remarkable plasticity to the façade.

1

tural precast concrete (Fig. 1.2.6),
the building is 48 stories tall and is
capped by a 212 ft (64.6 m) spire,
for a total height of 853 ft (260 m).
Floor-height double-window units,
weighing 3.5 ton (3.2 t) each, make
up half of the total precast pieces
used, with two variations for all corner units. It is just one of thousands
of buildings with a unique, distinctive façade that has been created
with architectural precast concrete.
The exterior precast concrete portion of the building in Fig. 1.2.7(a)
is over 900 ft (274.3 m) long. The
158 architectural precast concrete
panels are designed as an elliptical
curve, featuring a discontinuous
Fig. 1.2.6 TransAmerica Tower
San Francisco, California;
Architect: Johnson Fain and Pereira
Associates (formerly William L. Pereira
& Associates); Photo: Wayne Thom
photographer.

(a)

A building’s exterior can make a strong impression
on a visitor and enhance a company’s image. Take
the familiar landmark, the TransAmerica Building in
San Francisco, California. Clad entirely in architec(b)

Fig. 1.2.7(a) & (b) Center of Science & Industry, Columbus,
Ohio; Architect: Arata Isozaki & Associates, Design Architects:
NBBJ Architects Inc.; and Moody/Nolan, Architects of Record.

ARCHITECTURAL PRECAST CONCRETE

|

5

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

(a)

Fig. 1.2.8(a) & (b)
Merrill Lynch Facility
Englewood, Colorado;
Architect: Thompson, Ventulett Stainback &
Associates; Photos: Brian Gassel/TVS.

clothoid curve, which is a segment of a
spiral that curves in two directions. Each
quadrant of panels making up the façade
was placed along segments of six curves
to produce the elliptical shape. The wall
panels lean nearly 8 ft (2.4 m) into the
building along segments of two other circular curves (Fig. 1.2.7[b]). The architect
desired a smooth exterior surface. The result was a 20-in. (500 mm) deep wall panel comprised
of a 5-in. (125 mm) thick flange creating the exterior
shell and appearance. Each panel also contains two
vertical interior ribs to serve as panel stiffners. The 62ft-tall (18.9 m) curved panels not only clad the building
but also act as loadbearing members to support the
steel roof framing members and metal deck.
Two similar loadbearing total precast concrete office
buildings comprise a corporate campus. With their
long horizontal lines, strong vertical columns, and window frame details, the buildings are reminiscent of the
Prairie School style of architecture (Fig. 1.2.8[a]). The
rectangular forms maximize space planning efficiency
and accommodate the loadbearing precast concrete
structural system. Detailed with horizontal reveals, the
large, 30 ft wide x 40 ft high (9.1 x 12.2 m) red portals
replicate native Colorado red sandstone (Fig. 1.2.8[b]).
Buff-colored precast concrete tracery accents window
openings and forms pilasters at the third level, rising
into columns supporting the roof with its deep trellised sunshade. The buff precast concrete beneath the
windows features two finishes: a ribbed central section

6

|

ARCHITECTURAL PRECAST CONCRETE

(b)

between two acid-etched raised edges. The total precast concrete system also produces a single source of
responsibility for the structural system. This approach
lowered risk, streamlined communication, and made
on-site coordination easier for the design team, providing a significant value.
With its three parallel horizontally and vertically posttensioned curving walls constructed from three hundred forty-six 30-in.-thick (760 mm) off-white panels of
precast concrete that rise 57 to 88 ft (17.4 to 26.8 m)
above the nave, the church in Fig. 1.2.9(a) is a dramatic
and defining presence amid the suburb’s large, nondescript apartment buildings. The almost toppling walls
cantilever from the ground and all three walls are perfect segments of circles with the same radius. The shells
delineate three distinct spaces—the main sanctuary, the
weekday chapel, and the baptistery, each with its own
entrance. The secular community center is a concrete
structure with a rigidly rectilinear form (Fig. 1.2.9[b]).
The concrete is made with a white cement containing
photocatalytic particles of titanium dioxide that oxidize
organic and inorganic pollutants, so that the brightness

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.9(a) & (b)
Jubilee Church (Dives in Misericordia); Rome, Italy;
Architect: Richard Meier & Partners, Architects LLP; Italcementi,
Technical Sponsor; Photos: Gabriele Basilico.

1

(b)

(a)

ARCHITECTURAL PRECAST CONCRETE

|

7

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

and color will not degrade over time—resulting in selfcleaning concrete. In addition, the thermal mass of the
exposed interior concrete walls keep heat inside in the
winter and outside in the summer, reducing the energy
loads on the building system.
Various colors and textures of architectural precast
concrete are used to articulate the façade of the office building in Fig. 1.2.10. A rusticated stone-faced
first course is topped by exposed red granite aggregate
concrete forming the base of the building. The recessed
half-round columns, spandrels, and cornices are finished
in an acid-etched buff color simulating the appearance
of limestone. The repetition of precast concrete units
afforded the opportunity to make them more intricate
and created an elegant yet economical skin.
Fig. 1.2.10 Bannockburn Centre,
Bannockburn, Illinois;
Architect: Wright Architects, Ltd.;
Photos: Wright Heerema Architects.

(a)

The combination of color, shape, and texture showcases the ability of architectural precast concrete in Fig.
1.2.11(a) to meet the designer’s imaginative demands.
A large passageway frames the main entrance, involving architectural precast concrete that forms the columns and fascia beams. The project consists of four
main parts, namely a detention facility, the courthouse,
an addition to an existing police station, and the rotunda. Sixteen thin precast concrete shell units form the
rotunda’s 72 ft (22 m) diameter dome, bearing on a
Fig. 1.2.11(a) Boldly detailed loadbearing window-wall units.
Fig. 1.2.11(b)
Aurora Municipal Justice Center, Aurora, Colorado;
Architect: Skidmore, Owings & Merrill.

(b)

8

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

cast-in-place tension ring and a precast concrete compression ring. The rotunda is believed to be the first
precast concrete dome structure built in the United
States. The majority of the wall panels are loadbearing window-wall units, which are two stories high and
weigh 20,000 lb (9.1 t) each. They are boldly detailed
with bullnoses, cornices, and friezes, with a cream-colored appearance. Similarly, beveled reveals up to 10 in.
(250 mm) wide vertically score the centerline of each
panel to emphasize the play of sunlight and shadow
on the building’s monumental façade (Fig. 1.2.11[b]).

1

Fig. 1.2.12
W Hotel, San Francisco, California;
Architect: Hornberger + Worstell;
Photo: Hornberger + Worstell.

The hotel in Fig. 1.2.12 was designed for a prominent downtown corner site. It is flanked on one side
by boldly massed brick-faced precast concrete panels
on the Museum of Modern Art and has, as a nearby
neighbor, an art-deco designed, 1920s-style high-rise
building. To complement these prominent neighbors,
the architects generated a podium and tower design
with boldly massed pure forms articulated to reflect a
latent classical appearance. The tower was given deep
V-shaped scoring to emphasize its verticality. The 6in.-thick (150 mm) precast concrete panels (finished
in white concrete with black gray and buff aggregate)
were formed to look like blocks of granite used in
many of San Francisco’s civic structures. The appearance of stone detailing was enhanced by treating the
surface of the panels with two depths of sandblasting.
A medium sandblasting of the typical surface created
the look of a thermal finish, while a light sandblasting
of the back face of the reveals (3 in. wide x 2 in. deep
[75 x 50 mm]) at the lower levels was used to give the
impression of oversized mortar joints.
The art museum in Fig. 1.2.13 has become a landmark
symbol. The patterned façade has 1-in.-thick (25 mm)

Fig. 1.2.13 Museum of Modern Art
San Francisco, California;
Architect: Mario Botta, Design
Architect; Hellmuth, Obata &
Kassabaum, P.C., Architect of Record;
Photo: San Francisco Museum of
Modern Art/HOK/Perretti & Park
Pictures.

ARCHITECTURAL PRECAST CONCRETE

|

9

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.14 University of Illinois Molecular Biology Research Building, Chicago, Illinois; Architect: Goettsch Partners formerly Lohan
Associates; Photo: Jon Miller ©Hedrich Blessing.

brick on 9-in.-thick (225 mm) precast concrete panels.
Most panels measure 10 x 28.5 ft (3.0 x 8.7 m) and
contain between 1500 to 2300 bricks per panel.
The design challenge for the medical research building in Fig. 1.2.14 was based on the area’s extremely
eclectic architectural styles. Most of the buildings date
back to the 1920s and 1930s and feature a variety of
brick and precast concrete design motifs. There is no
dominant image to play off. The architects’ goal was
to create something that would be compatible with
this diverse group of buildings yet project a distinctive
campus image to visitors. The architects created a lot of
design interest in the articulation of the façade. In addition, by adding varying colors and finishes to the panels
based on their location, the designers were able to pick
up themes in nearby buildings without detracting from
the overall look. The result is that each façade is compatible with its surroundings and looks like it belongs in
the area. The base, a sandblasted granite aggregate finish, invites visual interest for pedestrians. The mid-levels
feature an acid-etched granite aggregate that provides
subtle, rich, red tones to integrate with the masonry
on existing buildings. The top features an acid-etched
white sand with a mild pink cast that is used primarily

10

|

ARCHITECTURAL PRECAST CONCRETE

on laboratory sections, conveying a crisp look. Ambertinted windows reinforce the color scheme.
The design challenge in Fig. 1.2.15 was to integrate
the multi-use complex into the small-scale principal retail district at the block’s edges and the adjacent art
deco building. The precast concrete and detailing at
the corner entrance speak to the tradition of the grand
era of late 19th-century department stores. From the
decorative urns at the roof, to the incised lettering, to
the terra cotta colored brackets and the strong cornices at the top of the second floor, this fineness of detail
was economically viable with precast concrete.
The project in Fig. 1.2.16(a) includes a 3-story shopping mall and an 11-story office building in the heart of
a major business district. The use of architectural precast
concrete panels to clad the buildings creates a distinctive
look, resembling rough-hewn stones. The panels also
comprise the creative geometric shapes that make the
complex stand out. The stone texture of the panels was
achieved by using rubber form liners, taken from natural
rocks in a basalt rock quarry (Fig. 1.2.16[b]). Detailed
engineering and intricate forms were needed to produce panels with deep reveals in false joints and interlocking lateral ends that eliminated any visible vertical

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

Fig. 1.2.15
Saks Majestic Square Complex
Charleston, South Carolina;
Architect: LS3P Associates, Ltd.;
Photo: LS3P Associates, Ltd.

(b)

(a)

Fig. 1.2.16(a) & (b)
Plaza Moliere Dos 22
Mexico City, Mexico;
Architect: Sordo Madaleno y
Asociados, S.C.;
Photo: Paul Citrón.

ARCHITECTURAL PRECAST CONCRETE

|

11

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

(b)

(a)

Fig. 1.2.17(a) & (b)
University Center of Chicago, Chicago, Illinois; Architect: Antunovich Associates and VOA Associates, Inc., Associate Architects;
Photos: Antunovich Associates.

12

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

Fig. 1.2.18 University of Iowa, Pappajohn Business Administration Building, Iowa City, Iowa;
Architect: Architecture Resources Cambridge, Inc., Design Architect; Neumann Monson, P.C., Architect of Record;
Photo: Nick Wheeler © Frances Loeb Library, Harvard Design School.

joints. The horizontal, vertical, and slanted corner panels
were cast in one piece, producing a true “corner stone”
appearance. The building is located within a high-risk
earthquake zone. The connections used a 1.25 in. (32
mm) slot as a provision for differential movement between any two adjacent stories and this was accomplished with 8-in.-long (200 mm) galvanized steel rods
that attached the precast concrete panels’ top inserts to
embedded plates in the structure.
About 1400 architectural precast concrete panels were
used to clad the 18-story, multi-university dormitory
project in Fig. 1.2.17(a). The design features rounded
corners and a detailed cornice, adding interest to the
massive structure. The panels feature an acid-etched finish on the upper levels and a retarded finish on the first
two floors, creating an appearance along the street that
fits with its neighbors (Fig. 1.2.17[b]). The extensive use
of the precast concrete allowed the exterior envelope to
be constructed during winter and inclement weather.

a harmonious rhythm of pilasters and windows, symmetry, and classical detail. The scale and feeling of the
exterior reflect the grace and serene neo-classicism of
nearby early 19th-century structures. The architectural precast concrete components for this project were
manufactured with a blend of white and gray cement,
a small amount of buff pigment, a light-colored limestone aggregate, and a natural buff sand. All of the
precast concrete panels were lightly acid-washed. The
precast concrete material’s extraordinary flexibility allowed considerable design freedom in the detailing of
the elements and the development of an aesthetic that
is sympathetic to the neighboring buildings. The precast
concrete creates a stepped, animated façade rather than
an unrelenting flat surface. Many of the precast concrete components required multiple casting sequences
due to the deep overhangs and multiple returns. The
precast concrete has an especially strong expression in
the rusticated base of the building, which conveys the
strength, solidity, and mass of the structure.

The façades of the four-story, block-long university
administration building in Fig. 1.2.18 are articulated by

For 85 years Oklahoma’s Capitol building remained
unfinished, its dome never built. The architect designed

ARCHITECTURAL PRECAST CONCRETE

|

13

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.19 Oklahoma State Capitol Dome, Oklahoma City, Oklahoma; Architect: Frankfurt-Short-Bruza Associates, P.C.;
Photo: Frankfurt-Short-Bruza Associates.

14

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

Fig. 1.2.20(a) & (b)
840 N. Lake Shore Drive, Chicago, Illinois;
Architect: Lucien Lagrange Architects;
Photos: Steinkamp Photography.

(b)

Corinthian columns and capitals, Greek pediments, ornate scrolled brackets, and highly detailed cupola. To
address concerns regarding the required maintenance
of the sealed precast concrete panel joints, precast
concrete ribs were placed over the vertical joints of the
dome panels. As an added precaution, all inside faces
of the precast concrete panels are accessible to help
prevent damage to the inner dome from undetectable
leaks.

(a)

both an external and inner dome based on the original
intent of the 1914 plans. The new outer dome features
architectural precast concrete and cast stone. The dome
is 80 ft (24.4 m) in diameter and rises 140 ft (42.7 m)
above the existing roof (Fig. 1.2.19[a] and [b]). The
greatest challenge for the precast concrete was to emulate the ornate detailing of the Capitol, which included

In a city renowned for its architectural heritage,
the distinctive 26-story condominium tower in Fig.
1.2.20(a), rising from a prominent lakefront site, is a
welcome addition to the skyline. Architectural precast
concrete panels clad the tower’s façade and help the
project integrate seamlessly into the neighborhood
through the articulation and warm buff color of the
material. The goal was to achieve the appearance of
an all-limestone-clad building, and this was enhanced
by the use of natural stone at the base, with precast
concrete panels used above the second floor. Typical
panels on the front elevation are 23 ft 3 in. long x 10
ft 8 in. high (7.10 x 3.28 m). Panels are 7 in. (175 mm)
thick with 11 in. (275 mm) returns at the windows (Fig.
1.2.20[b]). The cladding was designed for a 127 mile
per hour (203 km/hr) lateral wind load. The building’s
deep-punched windows, jointing, and cornice line
were achieved with precast concrete’s flexibility and
low cost. The designers also took advantage of precast
concrete’s fluidity in producing an elegantly curved
rotunda transition at the building’s most prominent
corner. The use of precast concrete panels minimized
both construction time and the staging area required
for installation, a key factor in an urban environment.

ARCHITECTURAL PRECAST CONCRETE

|

15

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

The precaster scheduled timely deliveries that could
be unloaded and transported via tower crane to their
connection location without stockpiling any pieces at
the site. The use of precast concrete also allowed the
window units to be anchored easily into the precast
concrete, eliminating the need for a metal strongback
system that would have been required to attach them
to stone.
The corporate center in Fig. 1.2.21 is a focal point
for a new development area. It is integrated into the
landscape with the use of architectural precast concrete panels in an innovative form and expression. The
pyramid-shaped building holds exhibition areas, showrooms, and a multiple-use auditorium. The cube-shaped
building is eight stories high, topped with a penthouse,
and is used as corporate office space. Waterfalls cascade onto the floor inside the circular openings of the
four walls, then flow into reflecting pools in the main
plaza. To match the surrounding landscape and the design concept of natural stone, a combination of white
and gray crushed marble coarse aggregate and sand
with white and gray cement was crucial in giving the
buildings the right color blend. The architect sought
texture and brightness for the desired reflection of light
by using a medium-deep surface texture achieved with
pneumatic chisel tools. The marble chips shone when
the skin was broken off the precast concrete panels.

Fig. 1.2.22 The Bushnell Center for the Performing Arts
Hartford, Connecticut;
Architect: Wilson Butler Lodge Inc.;
Photo: Robert Benson Photography.

The architect also requested that slight color
variation be randomly added to the panels
in order to attain the natural pyramid stone
effect.

Fig. 1.2.21 Calakmul Corporate Center, Mexico, D.F., Mexico;
Architect: Agustin Hernández Navarro; Photo: Agustin Hernández.

16

|

ARCHITECTURAL PRECAST CONCRETE

Figure 1.2.22 is a performing arts center
used for the presentation of Broadway shows,
ballet, opera, symphony concerts, and productions by regional arts groups as well as cultural arts–related educational programs. The
project, a 90,000 ft2 (8370 m2) addition to
an existing historic theater, owes much of its
aesthetic success to the use of contemporary
precast concrete elements that complement
the historic Georgian limestone façade of the
1929 building. The design of the entry pavilion
centered on the use of carefully detailed and
molded precast concrete columns and spandrels. The precast concrete panels allowed the
architects to create a light, open design with
spans impossible to create in stone.

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

With no courtroom construction in the area for more
than 40 years, new courts were needed to meet the
justice requirements of a large urban area with an
architectural design that reflects the look, character,
fundamental strength of the institution, environmental
characteristics of the site, and the progressive judicial
body. A 295,000 ft2 (27,406 m2), 10-story courthouse
was conceived with a rooftop helipad on an irregular site (Fig. 1.2.23). The courthouse contains eight
courtrooms, with space for six future courtrooms,
and judicial support departments. The court building
was massed with a blend of design elements by using
curved precast concrete panels that form the judicial
court block opposed by a contemporary insulated glass
curtain wall for the administrative and public areas.
The precast concrete panels with sandblast finishes
were selected for their substantial, low-maintenance,
and timeless appearance, along with the security characteristics. The main public entrance is enhanced by
a two-story atrium. The building is secure through
vehicle barriers provided by stepped hardscape.

1

Fig. 1.2.23
Los Angeles County Municipal Court
Los Angeles, California;
Architect: Mosakowski Lindsey Associates;
Photo: Benny Chan Fotoworks.

A highly visible, sloping site directly adjacent to a
small lake was the impetus for creating a classical style
office building that would use the water as a giant reflecting pool (Fig. 1.2.24). A high level of detail was
achieved through the use of sharply articulated panels
that create strong shadow effects. Two precast concrete finishes were chosen. Lightly sandblasted surfaces give the appearance of a buff-color limestone when
viewed from a distance. Heavy sandblasted portions of

Fig. 1.2.24 Four Lakepointe
Charlotte, North Carolina; Architect: Urban Design Group Inc.; Photo: Urban Design Group.

ARCHITECTURAL PRECAST CONCRETE

|

17

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.25 Logistical Systems Operations Center, Wright-Patterson Air Force Base, Dayton, Ohio; Architect: KZF Inc.;
Photo: KZF Design Inc.

the panels simulate thermal finish granite to complement polished, pink granite insets. Granite and marble
insets are separated from the precast concrete finishes
by a 3/4 in. (19 mm) reveal at all locations. This also provided a way to disguise panel joints and create more
shadows and surface detail. False joints are added at
mid-spandrel and mid-column heights to give the illusion of smaller pieces.

The façade of the building in Fig. 1.2.25 is constructed of lightly sandblasted architectural precast concrete
panels with a design that is a direct synthesis of the
simplified art deco style of the adjacent 1940s and
1960s buildings clad in precast concrete. Precast concrete planters with low ground cover serve as crash-resistant vehicle barriers for the windowless, single-story,
data processing area.
Fig. 1.2.26
One Balmoral Avenue
Toronto, Ontario, Canada;
Architect: Rafael + Bigauskas Architects;
Photo: Rafael + Bigauskas Architects.

18

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

Fig. 1.2.27(a) & (b)
Richard E. Lindner Athletics Center,
University of Cincinnati, Ohio;
Architect: Bernard Tschumi Architects,
Design Architect; Glaserworks, Architect
of Record; Photos: Bernard Tschumi
Architects.

(b)

(a)

The nine-story, mid-rise, U-shaped condominium
project accommodates a total of 137 units (Fig.
1.2.26). The building terraces downward to a
height of three stories, thus addressing the lowrise residential enclave behind. The nine-story
façade is broken down by a one-story rhythmic
sequence of repeated cladded columns, which
are capped by a continuous, heavily articulated
cornice band at the second floor. The columns
connect to this cornice band, emphasizing its support of the façade above. The elevation above the
cornice band is broken up by the integration of
punched windows and curved French balconies
that are framed through the use of detailed precast
concrete lintels and sills.
The boomerang-shaped, 236,000 ft2 (21,924 m2) athletic center is not only a fantastic work of art, but also
an amazing logistical accomplishment made possible
by architectural precast concrete (Fig. 1.2.27[a]). The
diagrid exoskeleton was conceived as a stiff structural
skin that could minimize the number of exterior columns while efficiently supporting clear spans over an
existing arena, mechanical room, service tunnel, and
loading dock. What resulted is an artfully conceived and
carefully executed integration of form and structure:
an extremely complex steel truss frame clad with 567
compound-curved, light gray, light sandblasted precast
concrete panels that actually wrap the individual steel
structural elements to provide deep window recesses
echoing the steel superstructure (Fig. 1.2.27 [b]). The

only straight lines in the precast concrete panels are in
the vertical plane or along the sides of the windows.
In the horizontal plane, there are six convex, two concave, and eight compound transition curves cast using
eight custom-constructed, all-steel forms—including
a one-of-a-kind adjustable form—that provided both
consistency and production efficiencies. In addition,
the structure appears perched atop several V-shaped
“pilotis” column covers that feature a helical warped
surface and vary in width throughout their height. The
helical steel frame and precast concrete panels with
recessed triangular window returns that penetrate into
and through the steel frame could only be designed
using data-laden 3-D software models and would have
been considered an almost unworkable design just a
few short years ago.
The dramatic, three-building library clad in architectural precast concrete is a functional work of art (Fig.

ARCHITECTURAL PRECAST CONCRETE

|

19

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

(a)

(b)

Fig. 1.2.28(a), (b) & (c)
Salt Lake City Public Library, Salt Lake City, Utah;
Architect: Moshe Safdie and Associates Inc., Design Architect;
VCBO Architecture, Architect of Record; Photos: Timothy Hursley.

1.2.28[a]). The largest and most demanding portion involved the 400-ft-long (120 m) sloping crescent building,
which used precast concrete to create the appearance
of a flowing, curved, leaning, and warping structure.
The crescent building climbs to a height of more than
90 ft (27 m) and at the midpoint of the inclined arc, the
inclination is 15 degrees from vertical (Fig. 1.2.28[b]).
It moves to embrace the main triangular library struc-

(c)

Fig. 1.2.29 Thomson Consumer Electronics Headquarters, Indianapolis, Indiana;
Architect: Boka Powell formerly Haldeman, Powell + Partners Consortium for Architecture Inc.; Michael Graves, Architect, Associate
Architect; Photo: BOKA Powell.

20

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

ture. Flowing through the skylit
space created by the triangle and
crescent buildings, the airy urban
room offers easy access to public
amenities (Fig. 1.2.28[c]). The sixstory rectangular structure contains
the administrative offices. All three
buildings are enclosed with, or incorporate the use of, acid-etched
architectural precast concrete
panels.
The administrative building in
Fig. 1.2.29 is organized as two office wings flanking a cubic central
pavilion housing a skylit cylindrical
atrium and grand stair. The designer modified the large scale of the
building’s long façades through the
use of a giant pattern. Integrally
colored green, gray, and terra cotta
precast concrete panels assembled
in a checkerboard pattern recall the
orderly landscape of the surrounding farm fields. The earth-colored
precast concrete central pavilion
provides a rich background against
which the bright yellow–glazed
brick portico reflects the brilliant
light. The choice of concrete panels was cost effective in addition to
providing a durable, low-maintenance material. Two other benefits
of precast concrete that helped
the construction schedule were its
ability to be installed in all types of
weather and the ease and speed
with which a precast concrete clad
building can be enclosed.
Architectural precast concrete panels, articulated in
an ashlar pattern, provide the design of the medical facility in Fig. 1.2.30 with the solidity and strength envisioned by the client. Precast concrete panels provided
an ideal solution to obtaining the warm buff coloration
produced from the process of chat-sawing limestone.
Precast concrete gave the designers control over color
and texture and provided flexibility in the sculptural expression. Half inch by half inch (13 mm x 13 mm) deep
reveals provided the necessary relief to create the ashlar
stone pattern used to reduce the scale of the large pan-

Fig. 1.2.30
Northwestern Memorial Hospital, Chicago, Illinois;
Architect: Ellerbe Becket Inc.; Hellmuth, Obata & Kassabaum,
P.C.; and VOA Associates Inc., Architects of Record;
Photo: Scott McDonald ©Hedrich Blessing.

els and simulate the texture of a natural stone surface.
A delicate, nearly smooth finish was produced by using a light sandblast that expressed almost no coarse
aggregate. A strong pilaster expression, enhancing the
vertical planes of the building, is also achieved through
the use of the precast concrete.

ARCHITECTURAL PRECAST CONCRETE

|

21

1

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

Fig. 1.2.31
Navy League Building, Arlington, Virginia; Architect: Page Southerland Page, L.L.P.; Photo: Doug Flory.

Architectural precast concrete was chosen for the
building cladding material in Fig. 1.2.31 for its inherent
design flexibility and sense of dignity and permanence.
It offered the ability to create a rich architectural vocabulary of façade elements establishing the appropriate
scale and image sought. Due to the extreme prominence
of this facility within a very active, energized community,
it was necessary to give the seven-story office building
a great deal of architectural detail and character. The
design team was admonished to avoid “flat” façades;
hence, the final design incorporates a significant level
of three-dimensionality with heavy vertical projections.

22

|

ARCHITECTURAL PRECAST CONCRETE

Three mixture designs were used with two finishes—
white precast concrete with embedded granite panels
and a light acid wash finish at the lower levels, and two
buff-colored mixtures with medium acid washes above.
Four spandrel panels have molded lettering showing the
building name and address. The building was designed
to achieve a Silver Rating from the U.S. Green Building
Council’s LEED Rating System.
Architectural precast concrete was selected to achieve
the classical look that the architect desired for the main
library in Fig. 1.2.32(a), which was a unique finish to

STATE-OF-THE-ART

1.2 Applications of Architectural Precast Concrete

1

(a)

(b)

Fig. 1.2.32(a) & (b)
Jacksonville Public Library
Jacksonville, Florida;
Architects: Robert A.M. Stern Architects; and
Rolland DelValle and Bradley;
Photos: Brian Griffis.

approximate the surface of coquina stone (appearance
is similar to travertine), a coarse-textured whitish limestone formed of broken shells and corals cemented together. In order to acquire this look, baking soda was
sprinkled in the bottom of the mold before the panel
was cast. Each rusticated panel has its own personality
with a wide variation in texture, veinage, and color,
but an average depth of surface voids was set to the
architect’s requirements (Fig. 1.2.32[b]). The resulting
appearance gives the building uniqueness.
Improvements in fabricating processes allow architectural precast concrete to be created in almost any
color, form, or texture—whatever is most aesthetically
pleasing. In addition, concrete’s moldability offers the

freedom to sculpt the structure’s façade in very imaginative ways. This ability to achieve totally customized
elements makes precast concrete different from any
other exterior cladding material. Precast concrete also
can be faced creatively with a wide variety of other
cladding materials. Architects are incorporating the
pleasing appearances of traditional cladding materials
such as dimensional stone, brick, tile, and even terra
cotta with the strength, versatility, and economy of
precast concrete.

ARCHITECTURAL PRECAST CONCRETE

|

23

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

Fig. 1.3.1
Purdue University Water Sculpture, West
Lafayette, Indiana; Robert Youngman, Sculptor;
Photo: Purdue University.

1.3 M
 ISCELLANEOUS USES OF
PRECAST CONCRETE
In addition to functioning as exterior and interior
wall units, architectural precast concrete finds expression in a wide variety of aesthetic and functional uses,
including:

The site design of a hospital includes three intaglio
art panels totaling 65 ft (19.8 m) in length (Fig. 1.3.2).
Set in a pebble garden sitting at ground level, the panels frame the entry plaza and provide a background
for the tide pool garden located behind the panel. A
prominent artist from California was commissioned for
the project and spent a week in the precaster’s plant

• Art and sculpture
• Columns, bollards, lighting standards, and fountains
• Planters, curbs, and paving slabs
• Towers
• Balconies
• Sound barriers and retaining walls
• Screens, fences, and handrails
• Street furniture and obelisks
• Ornamental work
• Signage
Artists have found precast concrete well suited for
expressions of boldness and strength in a variety of
shapes, forms, and textures, as illustrated in Fig. 1.3.1.

24

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 1.3.2
Hoag Hospital Sue and Bill Gross Pavilion
Newport Beach, California; Architect: TAYLOR
Tom Van Sant, Sculptor; Photo: Michael McLane and Courtesy of
TAYLOR.

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

1

(b)

(a)
(c)

Fig. 1.3.3(a), (b) & (c)
Metronome, One Union Square South, New York, New York;
Architect: Fredenburgh Wegierska-Mutin Architects; Kristin
Jones and Andrew Ginzel, Art Wall Designers; Photos: David
Sunberg/ESTO.

placing custom-designed foam impressions in the mold
to create the art. The foam was sandblasted away to
reveal a stunning California coastal shoreline–inspired
artwork within the sand-colored panels. The panels depict shorebirds along the water’s edge, which can be
seen from the site’s spectacular coastal bluff location
that overlooks the Pacific Ocean.

ARCHITECTURAL PRECAST CONCRETE

|

25

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

Fig. 1.3.4(a) & (b)
Kohl Center, University of Wisconsin, Nichols-Johnson Pavilion
Madison, Wisconsin;
Architect: Hellmuth, Obata & Kassabaum, P.C. and
Heinlein and Schrock (joint venture); Photos: Steve Brock.

(b)

(a)

The art wall sculpture serves as the main façade of
a 27-story, mixed-use residential-retail building (Fig.
1.3.3[a]). The art wall is 50 ft (15.2 m) wide by 100
ft (30.5 m) high with 36 concentric rings, with wave
spacing varying from 15 to 36 in. (380 to 910 mm),
and trough depths from 14 to 9 in. (355 to 225 mm).
The molds were handmade from full-sized, 3-D computer-generated patterns to ensure compliance with
the exacting requirements (Fig. 1.3.3[b]). Over 50,000
bricks in 8 different shapes were laid concentrically in
29 unique panels weighing from 8 to 22 tons (7.2 to 20
t). Precast concrete panels proved to be the best candidate by far for the design of the undulating pattern
of waves and rippling fluidity (Fig. 1.3.3[c]). Identical

26

|

ARCHITECTURAL PRECAST CONCRETE

bricks to those used in the art wall sculpture were used
for the rest of the building. The intent was to give the
appearance of the building being in motion.
The images of basketball players were incised into the
panel surface by placing a wooden pattern in the mold
(Fig. 1.3.4[a]). The master mold concept was used; the
bulkheads and side rails were moved to give the appearance of the players coming out of the ground (Fig.
1.3.4[b]). The outer surface was sandblasted and the
incised surface was left with a smooth as-cast finish.
Early 20th-century neoclassical commercial buildings
and mass-produced cast iron office buildings inspired
the design of the office building in Fig. 1.3.5. This

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

1

Fig. 1.3.5
1501 M Street, Washington, D.C.;
Architect: Hartman-Cox Architects;
Photo: Alan Karchmer.

ARCHITECTURAL PRECAST CONCRETE

|

27

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

curtain wall. The acid-washed
architectural precast concrete
column wall is self-supporting
and is tied back to the building’s
concrete structure with lateral
ties.
The aesthetic and functional
uses of precast concrete have
greatly expanded with the increased development of pedestrian malls and plazas. Precast
concrete artwork and functional sculpture, lighting standards, bollards, signage, and
fountains are frequently seen in
these areas. Planters and street
furniture are another application of precast concrete that
has gained importance with the
proliferation of malls and plazas. Street furniture can benefit
from the clean lines and variety
of textures possible with precast
concrete.

Fig. 1.3.6
Orlando International Air Traffic Control Tower, Orlando, Florida;
Architect: URS Corporation (Radian International);
Photo: Hensel Phelps Construction Co.

building’s exterior consists of two layers: a freestanding
screen of 175 two-story-high Doric columns and cornices placed in front of a finely detailed, painted metal

28

|

ARCHITECTURAL PRECAST CONCRETE

The design development of
air traffic control towers continues to evolve as taller facilities
are required to view the airport
surfaces of the nation’s larger
airports. Figure 1.3.6 is one of
the tallest towers in the United
States at 310 ft (94.5 m) from
cab floor to the first level. The
precast concrete panels of the
shaft are joined vertically with
grouted splice sleeves located on
each face of the panels. All panels were shimmed and grouted
at the horizontal panel-to-panel
joint. Panel-to-panel horizontal connections were also accomplished through the
use of mechanical splice sleeves. The entire exterior of
the tower (and both support buildings) were stained
with a two-color system to give the tower its signature
identity.
The structure in Fig. 1.3.7 serves as both a church bell
steeple and a cellular communications tower. Because
of the aesthetic requirements of the church, it was

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

1

Fig. 1.3.7
All Saints Catholic Church Bell Steeple
Dunwoody, Georgia;
Architect: Slater-Paull & Associates Inc.;
Photo: Slater-Paull & Associates Inc.

ARCHITECTURAL PRECAST CONCRETE

|

29

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

legs gives the steeple a dramatic
and monumental appearance. The
three-sided arched members that
connect the steeple legs together
are cast as one unit. Each of these
three legs consists of five precast
concrete members with a base
spread of 65 ft (19.8 m). Two horizontal pieces and two arched segments are placed at intermediate
heights, for a total height of 129
ft (39.3 m).
Some major advantages of using
precast concrete for the bell tower
in Fig. 1.3.8 were ease of construction of the complex structural design and availability of complex
shapes and finish selection. Another
major advantage is that the precast
concrete walls carry the weight of
the tower down into the concrete
foundation, which rests directly
on bedrock, making it possible to
have a 160 ft (48.8 m) freestanding
carillon tower. The structural frame
is made up of three structural elements working simultaneously in
conjunction with one another. The
structural steel frame, eight large
steel compression rings, and posttensioned precast concrete panels
are each tied to one another in alternating connection sequences.

Fig. 1.3.8
Millennium Carillon, Naperville, Illinois; Architect: Charles
Vincent George Design Group Inc.; Photo: Charles Vincent George
Design Group Inc.

important to hide the structure’s dual function. This
concern became the prime motivation for choosing
precast concrete. All transmission cables and lightning
protection grounding systems are concealed inside of
PVC conduit, which is cast into the steeple legs. The 40
ft (12.2 m) radius curve defining the precast concrete

30

|

ARCHITECTURAL PRECAST CONCRETE

The two most dramatic features
of the garden crypt complex are the
50-ft-tall (15.2 m) precast concrete
entrance tower, with its striking
terra cotta artwork (Fig. 1.3.9[a]),
and the 75-ft-long (22.9 m) black
granite fountain which serves as a central axis through
the building and delivers the sound of falling water
throughout the entire complex (Fig. 1.3.9[b]). The
tower consists of 16 precast concrete pieces, including
concave and convex radius pieces and gable-roofed,
pediment-shaped pieces. The project also includes
thirty-four 1 ft 8 in. (0.5 m) diameter cylindrical columns with curved spandrels in the central courtyard.
The finishes of the architectural precast concrete are a
medium sandblast and acid-etched finish.

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

1

Fig. 1.3.9(a) & (b)
All Saints Cemetery – Immaculate Heart of
Mary Mausoleum
Des Plaines, Illinois;
Architect: Mehus Studio, Ltd.;
Photos: Craig Dugan ©Hedrich Blessing.

ARCHITECTURAL PRECAST CONCRETE

|

31

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

Most of the balustrades in Fig. 1.3.10
were cast in two-piece fiberglass forms
that opened to allow stripping of the precast concrete. Some balusters were cast
individually and then joined with the top
and bottom rails in the field to form the
desired length of balustrade. The stairs
are also precast concrete.
The main feature of the parking structure in Fig. 1.3.11 is a 50-ft-tall (15.2 m)
bookshelf on one side. The wall panels
are curved on the outside face to resemble book spines sitting on a library shelf.
A photographic vinyl film is attached to
a light gauge metal sheet, which in turn
is attached to the precast concrete. The
structure’s grand stairs are also made of
precast concrete designed to look like a
Fig. 1.3.10
Millennium Park, Chicago, Illinois;
Architect: Skidmore, Owings & Merrill.

stack of books. Form liners were used to create the
look of book pages.
Sunscreens may be loadbearing, wall-supporting, or
part of cladding. They may also be freestanding when
used as dividers or fencing. Architectural sunscreens
and large solid wall panels may be used to reinforce
a strong design statement. Screens are often used to
decoratively shield the space from sunlight or to block
specific areas from public view, see page 397. They
may also serve to renovate older buildings.
Barrier walls have become popular in recent years because of the growing need to control excessive noise
pollution generated by busy airports and highways.
Using a variety of finishes and textures on the outer
surface, the barriers may be designed to blend in with
the adjacent neighborhood.
More than 200 sculptured precast concrete sound
barriers, Fig. 1.3.12, depicting wild life images from
the surrounding landscape, were installed along a
Fig. 1.3.11
Library District Parking Garage
Kansas City, Missouri;
Architect: BNIM and 360 Architecture (joint venture);
Photo: BNIM/360.

32

|

ARCHITECTURAL PRECAST CONCRETE

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

1

Fig. 1.3.12
Maryland Interstate Highway 216
Howard County, Maryland;
Sculptor: Creative Design Resolutions, Inc.;
Photos: Creative Design Resolutions, Inc.

ARCHITECTURAL PRECAST CONCRETE

|

33

1

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete

Fig. 1.3.13
Rotunda at Carnegie Center, Charlotte, North Carolina;
Architect: LS3P Associates Ltd. formerly Clark Tribble Harris and
Li Architects, P.A.;
Photo: Rick Alexander & Associates, Inc.

busy, one mile (1.6 km) stretch of interstate highway.
The modular sound barriers were designed to represent indigenous flora and fauna, including trees, birds,
water, and landscapes.

34

|

ARCHITECTURAL PRECAST CONCRETE

In Fig. 1.3.13, a 500-ft-long (152 m) and 20-ft-high
(6.1 m) precast concrete waterwall conceals the twolevel parking structure below the building. This feature
wall consists of 32 panels of gray precast concrete

STATE-OF-THE-ART

1.3 Miscellaneous Uses of Precast Concrete / 1.4 Benefits and Advantages of Architectural Precast Concrete

1

ect develops, the design team and
precaster can discuss panelization,
types of finishes, shapes, repetitive
use of efficient and economical building modules, structural systems, delivery schedule, erection procedures,
and sequencing. This time spent in
development will pay off in accelerated construction time and significant
cost savings. As a result, it is essential
that the designer work with the local
architectural precast concrete producer in the early design stages and
throughout the development of contract documents. Properly implemented, an early and continuing dialogue
between the designers and the precaster will ensure optimum product
quality and appearance at a minimum
installed-construction cost.
Fig. 1.3.14
Millennium Park, Chicago, Illinois;
Architect: Skidmore, Owings & Merrill.

between vertical columns of cream precast concrete.
Each gray panel has a decorative relief design cast in
the abstract form of a tree with the stylized “limbs”
on each side of the “trunk” continuously sprayed by
individual horizontal water jets concealed in the cream
columns.
As a replacement for terra cotta or stone ornamentation, precast concrete is an ideal material. Impressions
can be taken of the ornaments and molds made to cast
the replacements in concrete with highly refined, intricately detailed relief patterns. With sufficient repetition
to amortize initial mold-making costs, ornamental precast concrete can be economically feasible.
Precast concrete with incised lettering can be used to
produce identifying signs for parks or office development (Fig. 1.3.14).

1.4 B
 ENEFITS AND ADVANTAGES OF
ARCHITECTURAL PRECAST CONCRETE
When the design team works with the precaster from
the outset of a project, it is more likely that the full benefits of precast concrete will be realized. As the proj-

As the first and often longest-lasting impression, the
exterior of a building is its signature. But a building
envelope’s materials are more than a visual application.
Aesthetics, function, and cost play a role in achieving
a successful project. Architectural precast concrete not
only offers design freedom of architectural expression
with visually interesting shapes that are functional in
application, it contributes to durability, sustainability,
energy efficiency, and improved occupant comfort and
safety. At the same time, the plasticity of concrete allows the designer to achieve a high level of detail in the
profile, scale, and character of a building that cannot be
matched by other materials due to costs.
Durability: Architectural precast concrete panels provides proven long-term durability. It provides a façade
that is exceptionally resistant to impact, corrosion, weathering, abrasion, and other ravages of time, making it
virtually maintenance-free and resulting in preservation
of the building’s original look. The high cement contents
and low water-cement ratios used in the precasting process, combined with proper compaction and curing in a
controlled factory environment, ensure a dense, highly
durable concrete. A low water-cement ratio concrete
has been proven to resist weathering and corrosion. Airentrainment is used to improve freezing and thawing
resistance, particularly in severe environments.
Aesthetics: Architects find that precast concrete panels provide an unlimited vocabulary that allows design

ARCHITECTURAL PRECAST CONCRETE

|

35

1

STATE-OF-THE-ART

1.4 Benefits and Advantages of Architectural Precast Concrete

concepts to be executed in a broad range of architectural styles, shapes, and sizes. The material offers limitless
potential for developing and manipulating mass, color,
form, texture, and detail to obtain simple, clean shapes
that enhance an image of strength with aesthetic beauty. Different aggregates, color tones, textures, and patterns can be designed for each building in a complex
to differentiate them. The initial plasticity of precast
concrete makes it responsive to the designer’s creative
needs. Precast concrete mold-building techniques allow
designers to enhance a building’s visual interest through
elements such as ribs, bullnoses, reveals, chamfers, or
textures. Designers can economically incorporate details
such as cornices, quoins, arches, and decorative relief
panels. In addition to these benefits, the ability to manipulate color, form, and texture make precast concrete
an excellent material to consider in situations where
the relationship of a building to its existing context is
an important design consideration. Precast concrete can
be designed to harmonize with and complement other
materials. Natural stone, brick, tile, or terra cotta can be
cast into panels allowing designers even more choices
for panel finishes. Precast concrete can also replicate
the color and finish of a wide variety of costly facing
materials.
Commitment to Quality: Architectural precast concrete units produced by PCI-Certified plants are manufactured under strict, factory-controlled conditions to
an agreed production schedule ensuring a uniformly
high-quality façade in the desired shapes, colors, and
textures. To become PCI Certified, producers must satisfy an array of production, administrative, and organizational procedures along with close tolerances unique

to precast concrete. To maintain certification, every PCI
member must undergo stringent unannounced inspections each year by independent auditors. It is strongly
recommended that PCI Certification be a requirement
for the award of the precast contract to help maximize
the quality of the finished product. This requirement
may minimize the need for continuous inspections by
the owner.
Life Cycle Cost: The life cycle cost of the structure is
another area where architectural precast concrete exhibits superior performance. A precast concrete façade
can be designed to match the intended life of a building
with minimal maintenance, providing substantial longterm savings. Precast concrete panels present a durable,
aesthetically pleasing exterior surface that is virtually air
and watertight and does not require painting. This helps
the building remain in first-class condition ensuring its
desirability for future tenants or owners.
Initial Cost: Precast concrete’s speed of erection and
ability to be cast and erected year-round aids the entire
construction team. Because the casting process does not
rely on other critical-path activities to begin, units can be
produced as soon as drawings are approved, ensuring
units are ready for erection as soon as foundation work
or supporting structure is completed (Fig. 1.4.1). These
advantages allow the building’s shell, whether loadbearing or cladding, to be enclosed quickly. This in turn allows interior trades to begin work earlier and compresses
the overall building schedule. Faster completion reduces
interim financing and construction management costs,
results in earlier cash flows, and produces other economic benefits. This ultimately lowers the building’s

Fig. 1.4.1 Architectural Precast Concrete Cladding Project Schedule
SCHEDULE
Week #
Event/Task

1

2

3

4

5

6

7

8

9

Commitment to Precaster
Sample Approval
Design Development
Shop Drawings
Shape & Layout Drawings
Approval
Erection Drawings
Approval
Production Drawings
Mold Fabrication
Panel Fabrication
Panel Installation

Note: Schedule durations are for illustration only.

36

|

ARCHITECTURAL PRECAST CONCRETE

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43

STATE-OF-THE-ART

1.4 Benefits and Advantages of Architectural Precast Concrete

long-term overall cost and can make the use of precast
concrete more economical than other façade materials.
Loadbearing panels can further reduce framing costs
by providing a column-free perimeter. Depending on
the floor plan, there also is a potential for reducing the
number and/or size of interior columns, thereby aiding
layout flexibility. This results in a more efficient and less
costly construction. Cost savings of loadbearing panels
are greatest for low to mid-rise structures of three to ten
stories with a large ratio of wall-to-floor area.
Energy Efficiency: Precast concrete panels can be designed to provide a high degree of energy efficiency for
the buildings they enclose. Recessed window walls, vertical fins, and various other sculptured shapes facilitate
the design of many types of shading devices for window
areas to reduce glare and solar gain. This provides economies in the cost of the air-conditioning system by reducing thermal load. Specific thermal characteristics of
the wall can be designed for each face of the structure
to suit its sun orientation. To reduce heating and cooling costs, precast concrete walls may either have insulation field applied to their backs, or incorporated at the
plant to create sandwich wall panels. The thermal mass
inertia of concrete, which is recognized in American
Society of Heating, Refrigerating and Air-Conditioning
Engineers (ASHRAE) standards, also reduces peak heating and cooling loads, thus saving energy year-round
by reducing large daily temperature swings. In addition,
precast concrete construction allows minimal air infiltration or exfiltration, reducing the potential for moisture
problems due to moist air migrating into a wall and
building.
Other Inherent Benefits: Architectural precast concrete is non-combustible with inherent fire resistant capabilities, creating a safe building envelope that helps
protect personnel, equipment, and the building itself.
That in turn reduces insurance rates. It also eliminates
the need and cost of additional fireproofing measures, except when used as cladding on structural steel
frames.
Environmental Impact: The inherent sound attenuation properties, due to precast concrete’s mass, provide
an economical acoustical barrier to exterior and interior
noise penetration. These precast concrete panel attributes enhance their inherent cost effectiveness. The lifesafety and tenant benefits provide a potent marketing
asset when attracting long-term occupants.
Precast concrete is an environmentally sound material,

1

produced from natural materials. No toxic substances
are generated in use. Precast concrete uses by-products
from other industries that previously were landfill items.
Steel reinforcing bars are made from recycled automobile bodies and other recycled steel parts. Admixtures
used to control flowability and increase durability use
by-products from paper, aluminum, coal, and steel
plants in their manufacture. Also, the production energy consumption of the concrete is quite small. Precast
concrete construction allows minimal air infiltration;
thermal mass of concrete delays internal temperature
changes and reduces peak heating and cooling loads;
and sculptured shapes facilitate the design of shading
devices for window areas.
Precast concrete has the added quality of reflecting
heat as well as light, thus reducing the “heat island”
effect and higher temperatures common in urban areas. The resulting lower overall temperatures can make
a difference in the amount of electricity consumed in air
conditioning and can reduce smog formation, potentially improving air quality in urban areas.
Precast concrete wall panels can be reused when buildings are expanded. Non-loadbearing panels on the end
are simply disconnected from the framing and additional panels and framing are added on each side. With the
new addition in place, the end panels can be replaced.
Concrete measures up well in regard to sustainability. Precast concrete strikes a perfect balance between
meeting today’s needs and preserving natural resources
for tomorrow.
Single-Source Provider: Precast concrete cladding
panels provide a one-source solution for supplying the
exterior wall system. When precast concrete structural
floors along with loadbearing panels are specified, the
complete building shell can be supplied by one certified producer. This single-source solution ensures that
the complete responsibility and accuracy for satisfying the design specifications rests with only one supplier. The precaster also is responsible for coordinating
manufacturing and constructability issues, reducing the
number of subcontractors, and minimizing cost. Also,
the producer’s staff of engineers are available to assist
the design team from design conception through completed project.
Supplier Assistance: PCI Certified precasters can offer detailed expertise that allows the development of
design techniques, engineering innovations, and sched-

ARCHITECTURAL PRECAST CONCRETE

|

37

1

STATE-OF-THE-ART

1.4 Benefits and Advantages of Architectural Precast Concrete / 1.5.1 Plant Certification Program

uling improvements that save time and money from conceptual design to project completion. To maximize these
benefits, the design team should interact and partner
with the precaster early in the project’s development
stage. This ensures that each element is as cost effective
as possible, taking full advantage of precast concrete’s
inherent performance characteristics. The result will be a
functionally efficient and aesthetically pleasing structure
that meets or exceeds the project’s needs.

1.5 Q
 UALITY ASSURANCE AND
CERTIFICATION PROGRAMS
Quality assurance and plant certification are important items in the prefabrication process. This program
responds to an ever-increasing demand from the marketplace for quality products and services.
The owner or architect must be confident that materials, methods, products, and the producer’s quality control prodedures meet the requirements of the project.
This assurance is available by requiring in the project
specification that:
1. The precaster facility be certified by the PCI Plant
Certification Program;
2. The precaster have personnel certified in the appropriate levels of the PCI Plant Quality Personnel
Certification Program; and
3. The precast concrete erector be certified by the PCI
Field Certification Program or the precaster have a
qualified person to oversee the work of the erector.
Certification of precast concrete production facilities
means that an independent inspection body has confirmed the plant has the capability to produce quality
products and the in-house quality control system functions efficiently.

1.5.1 P
 lant Certification Program
Producers registered under the PCI Plant Certification
Program in Product Group A have demonstrated that
their processes for production and quality assurance
meet or exceed industry-wide standards based on the
PCI Manual for Quality Control for Plants and Production
of Architectural Precast Concrete Products (MNL-117).
Product Group A has two categories: A1 for major,
primary architectural panels and products; and AT for
miscellaneous architectural trim elements. Plants must

38

|

ARCHITECTURAL PRECAST CONCRETE

maintain a comprehensive, documented, and approved
quality system that is present in every aspect of their business. Each certified plant conducts a formal quality control program with a trained, permanent quality control
staff. The staff must meet the current requirements of
the PCI Quality Control Personnel Certification Program.
Conformance to the nationally accepted requirements
is determined by a minimum of two audits per year.
All audits are unannounced and most are two days in
length. The audits are conducted by specially trained
personnel employed by a national structural consulting
engineering firm under contract to PCI and accredited
by International Accreditation Service Inc. (IAS).
Audits cover all phases of production including shop
drawings, materials, production methods, product handling and storage, appearance, testing, record keeping,
quality control, personnel training, and safety practices.
Failure to maintain a production plant at or above required standards results in loss of certification. A current listing of plants holding certification is published a
minimum of four times each year and may be obtained
by calling PCI or by visiting the PCI website at www.pci.
org.
Care should be exercised in evaluating the effectiveness of other certification programs. Criteria should
include:
1. Unannounced audits. They are fundamental to the
program.
2. The auditor should be recognized as experienced in
the field of precast and prestressed concrete.
3. The auditor should have particular experience with
the products and production methods involved.
4. The auditor should be independent and not hired
by the general contractor/construction manager or
precaster.
5. The program should be based on the industry-approved and nationally recognized quality standards.
6. The auditor should view the entire fabrication
cycle.
7. The program should be executed by a single auditing agency that ensures uniformity for all-size companies throughout the United States.
8. The program and auditing agency should be recognized by major public and private agencies and
organizations.

STATE-OF-THE-ART

1.5.2 Plant Quality Personnel Certification / 1.6 Definitions

1.5.2 P
 lant Quality Personnel
Certification
Conducting an effective quality control program requires knowledgeable and motivated testing and inspection personnel. Each must understand quality basics, the necessity for quality control, how products are
manufactured, and precisely how to conduct tests and
inspections. PCI has been training quality control personnel since 1974. There are three levels of Plant Quality
Personnel Certification.
Plant Quality Personnel Certification, Level I
requires a basic level of understanding of the many quality control issues normally encountered in a precast concrete plant. It also requires current certification by the
American Concrete Institute (ACI) Concrete Field Testing
Technician Program, Grade 1. A candidate must have at
least six months of industry experience.
Plant Quality Personnel Certification, Level II
requires Level I as a prerequisite. Other requirements for
Level II include demonstration of a greater level of knowledge of the topics for Level I, as well as at least one year
of industry experience. Certification at Levels I and II is
accomplished by passing a written examination.
Plant Quality Personnel Certification, Level III provides significant instruction in concrete materials and
technology. Certification at this level requires Level II as
a prerequisite. The candidate must have two years of
industry experience or equivalent.

1.5.3 F ield Certification Program

S1 – Simple Structural Systems (horizontal decking
members, single-lift walls)
S2 – Complex Structural Systems (category S1 plus all
other structural products, including loadbearing architectural units)
Under the PCI Field Certification Program, field audits
of erecting crews are conducted by PCI-Certified Field
Auditors (CFAs) and audits of organizational administrative controls related to erection projects are conducted
by PCI-Certified Company Auditors. The auditing criteria are based on the industry’s quality, procedural, and
safety standards as presented in PCI Erectors Manual
– Standards and Guidelines for the Erection of Precast
Concrete Products (MNL-127); and PCI Erection Safety
for Precast and Prestressed Concrete (MNL-132). Audits
are conducted semi-annually on all of the erector’s primary crews and cover all phases of the erection process including pre-construction planning, practices and
procedures, equipment, safety, erection tolerances, and
quality control. At least one of the two annual field audits for each crew must be performed by a CFA who
works for a company independent of the erectors. In
addition, an annual company audit must be passed
certifying the company’s managerial and administrative
ability to meet the required standards. Failure to maintain work at or above required standards results in mandatory loss of PCI Qualified or Certified Erector status.
A current list of PCI Qualified and Certified Erectors is
published a minimum of four times each year or may be
obtained by visiting the PCI website at www.pci.org.

The Field Certification Program extends PCI quality assurance from fabrication (plant certification) to field installation. It has been found that participation in the program is
highly beneficial in improving the quality of the installation
and the safety and efficiency of the installation crews.

1.6 D
 EFINITIONS

The PCI Field Certification Program confirms the capability
of producer-erectors and independent erectors of precast
concrete structures to handle and install precast concrete
units in compliance with established industry standards.
Specifying a PCI Qualified or Certified Erector or precaster
with Certified Field Auditors (CFA) ensures the project specifier and owner that the erector has met the rigorous requirements of the PCI Field Certification Program.

Air-Entraining admixture is a chemical added to the
concrete for the purpose of providing minute bubbles of
air to the concrete during mixing to improve the durability of concrete exposed to cyclical freezing and thawing
in the presence of moisture.

An erector may be qualified or certified in up to three
categories:
A–A
 rchitectural Systems (non-loadbearing cladding)

1

Admixture is a material other than water, aggregates,
or cement used as an ingredient of concrete or grout
to impart special characteristics. These are usually employed in very small amounts.

Approval is an action with respect to shop drawings,
samples, and other data that the general contractor
or construction manager is required to submit. When
used in this context, approval is only for general conformance with the design requirements and compliance
with the information given in the contract documents.

ARCHITECTURAL PRECAST CONCRETE

|

39

1

STATE-OF-THE-ART
1.6 Definitions

Such action does not extend to means, methods, techniques, sequences, or procedures of construction, or
to safety precautions and programs incident thereto,
unless specifically required in the contract documents.
Architectural precast concrete refers to any precast
concrete unit with a specified standard of uniform appearance, surface details, color, and texture. It is either
a special or occasionally standard shape that through
application or finish, shape, color, or texture contributes to the architectural form and finished effect of
the structure; units may be part of the structural frame
or non-structural cladding and may be conventionally
reinforced or prestressed.
Backup mixture is the concrete mixture cast into the
mold after the face mixture has been placed and consolidated. It is a less expensive mixture than the face
mixture.
Bond breaker is a substance placed on a material to
prevent it from bonding to the concrete, or between a
face material, such as natural stone, and the concrete
backup.

Crazing is a network of fine cracks in random directions breaking the exposed face of a panel into areas
from 1/4 to 3 in. (6 to 75 mm) across.
Creep (or plastic flow) is the time-dependent deformation of steel or concrete due to sustained load.
Dap is the blocked out section at the support end of a
beam or floor and roof member.
Design (as a transitive verb) The process of applying the principles of structural mechanics and materials science, and interpreting code regulations to determine the geometry, composition, and arrangement of
members and their connections in order to establish
the composition and configuration of a structure.

Bugholes are small holes on formed concrete surfaces
created by entrapped air or water bubbles.
Bulkhead is a partition in formwork blocking fresh
concrete from a section of the mold or the end of a
mold that establishes the length of a precast concrete
unit.

Designer (prime consultant) is the architect, engineer,
or other professional responsible for the design of the
building or structure of which the precast concrete
forms a part.

Cladding (non-loadbearing panel) is a wall unit that
resists wind, seismic, or blast loads and its own weight
and is attached to the structural frame. In many applications, it supports glazing systems.

Draft is the slope of concrete surface in relation to
the direction in which the precast concrete element is
withdrawn from the mold; it is provided to facilitate
stripping with a minimum of mold breakdown.

Clearance is the interface space (distance) between two
items. Normally, it is specified to allow for product and
erection tolerances and for anticipated movement.

Drift is the lateral deflection of one level at its center
of mass at and above that level. It is the difference
in predicted movement of the structure between two
adjacent stories under lateral loads.

Construction Manager (CM) is a person or firm engaged by the owner to manage and administer the
construction.
Contract documents are the design drawings and specifications, as well as general and supplementary conditions

|

Cover is the distance between the surface of the reinforcement and the nearest concrete surface.

Design team is persons or firms engaged by the owner
or owner’s representative to perform the architectural
and structural design for the structure and/or to provide
services during the construction process.

Connections are a structural assembly or component
that transfers forces from one precast concrete member to another, or from one precast concrete member
to another type of structural member.

40

and addenda, that define the construction and the terms
and conditions for performing the work. These documents are incorporated by reference into the contract.

ARCHITECTURAL PRECAST CONCRETE

Engineer of record (EOR) is the registered professional engineer (or architect) who is responsible for
developing the design drawings and specifications
in such a manner as to meet the applicable requirements of governing state laws and of local building
authorties. The EOR is commonly identified by the professional engineer’s seal on the design drawings and
specifications.
Envelope mold is a box mold where all sides remain

STATE-OF-THE-ART

1.6 Definitions

in place during the entire casting and stripping cycle.
Erector is usually the subcontractor who erects the
precast concrete components at the site. The general
contractor may also be the erector.
Exposed aggregate concrete is concrete with the
aggregates exposed by surface treatment. Different
degrees of exposure are defined as follows:
L ight exposure — Only the surface skin of cement
and sand is removed, just sufficiently to expose the
edges of the closest coarse aggregate.
	Medium exposure — A further removal of cement
and sand to cause the coarse aggregate to visually
appear approximately equal in area to the matrix.
	Deep exposure — Cement and sand have been
removed from the surface so that the coarse aggregate becomes the major surface feature.
Face mixture is the concrete at the exposed face of
a precast concrete unit used for specific appearance
reasons.
False joint is scoring on the face of a precast concrete
unit; it is used for aesthetic or weathering purposes
and normally made to simulate an actual joint (see also
Reveal).
Form see Mold.
Gap-graded aggregate is a mixture with one or more
normal aggregate sizes eliminated and/or with a heavier concentration of certain aggregate sizes over and
above standard gradation limits. It is used to obtain a
specific, more uniform exposed aggregate finish.
General Contractor (GC) is a person or firm engaged
by the owner to construct all or part of the project.
The general contractor supervises the work of its
subcontractors and coordinates the work with other
contractors.
Hardware is a collective term applied to items used in
connecting precast concrete units or attaching or accommodating adjacent materials or equipment. Hardware is normally divided into five categories:
	Field or contractor’s hardware—Items to be
placed on or in the structure in order to receive the
precast concrete units, such as anchor bolts, angles,
or plates with suitable anchors.

1

	Plant hardware—Items to be embedded in the
precast concrete units themselves, either for connections and precast concrete erector’s work, or for
other trades.
	Field installed pre-erection hardware—Miscellaneous loose steel pre-welded or pre-bolted to the
structure.
	Erection hardware—All loose hardware necessary
for the installation of the precast concrete units.
	Accessory hardware—Items to be cast into the
precast concrete units, designed and supplied by
the trade requiring them.
Homogeneous mixture is a uniform concrete mixture used throughout a precast concrete element.
Inserts are connecting or handling devices cast into
precast concrete units.
Loadbearing precast concrete units are those precast concrete units that form an integral part of the
structure and resist and transfer loads applied from
other elements. Therefore, a loadbearing member
cannot be removed without affecting the strength or
stability of the structure.
Master mold is a mold that allows a maximum number of casts per project. Units cast in such molds need
not be identical provided the changes in the units
can be accomplished simply as pre-engineered mold
modifications.
Matrix is the portion of the concrete mixture containing only the cement and fine aggregate (sand); it binds
the coarse aggregate.
Mold is the container or surface against which fresh
concrete is cast to give it a desired shape; it is sometimes used interchangeably with “form.” (The term is
used in this manual for custom-made forms for specific
projects while the term “form” is associated with standard forms or forms of standard cross-section.)
Optimum quality is the level of quality (in terms of
appearance, strength, and durability) that is appropriate for the specific product, its particular application,
and its expected performance requirements. Quality
also refers to the totality of features and characteristics
of a product that bear on its ability to satisfy stated
needs. Realistic cost estimates for producing the pre-

ARCHITECTURAL PRECAST CONCRETE

|

41

1

STATE-OF-THE-ART
1.6 Definitions

cast concrete units within stated tolerances are factors
which must be considered in determining this level.
Non-Loadbearing precast concrete units see
Cladding.
Precaster is the firm that manufactures the precast
concrete components.
Precast engineer is the person or firm who designs
precast concrete members for specified loads and who
may also direct the preparation of the shop drawings.
The precast engineer may be employed by the precaster or be an independent person or firm to whom the
precaster subcontracts the work.
Prestressed concrete is concrete in which permanent
internal stresses have been induced by forces caused
by tensioned steel. This may be accomplished by:
	Pretensioning—The method of prestressing in
which the tendons (prestressing steel) are tensioned
(elongated) and then anchored while the concrete
in the member is cast around the tendons, and released when the concrete is strong enough to receive the forces from the tendon through bond.
 ost-tensioning—The method of prestressing in
P
which the tendons (prestressing steel) are kept from
bonding to the fresh (wet) concrete, then elongated
and anchored directly against the hardened concrete, imparting stresses through end bearing.
Quality assurance (QA) is all those planned or systematic actions necessary to ensure that the final product or service will satisfy given requirements for quality
and performance of the intended function. Typically,
the quality assurance effort will focus on the requirements of the overall project, thus identifying the quality control requirements for member fabrication.

|

Return is a projection of like cross-section that is 90
degrees to, or splayed from, the main face or plane
of view.
Reveal is a groove (rustication) in a panel face generally used to create a desired architectural effect. It is also
the projection of the coarse aggregate in an exposed
aggregate finish from the matrix after exposure.
Rustication see Reveal.
Samples are a group of units, or portion of material,
taken from a larger collection of units or quantity of
material, which serves to provide information that can
be used as a basis for action on the larger quantity or
on the production process; the term is also used in the
sense of a sample of observations.
	Design reference samples are usually small specimens, 12 x 12 in. (300 x 300 mm), made by the
precast concrete plant laboratory to provide the
designer with early conceptual ideas of color and
texture.
	Bid reference samples are normally small specimens, 12 x 12 in. (300 x 300 mm), made by a producer to show the designer what can be made locally and is used as a basis for the producer’s bid.
	Selection samples are larger than bid reference
samples and are made by the successful producer
before casting any units; they become the basis
for accepting the appearance of finishes. (Full-size
production run elements are then approved and become the final standard for acceptance.)
Sandwich wall panel is a wall panel consisting of two
layers (wythes) of concrete fully or partly separated by
a layer of insulation.

Quality control (QC) is those planned actions that
provide a means to measure and control the characteristics of members and materials to predetermined
quantitative criteria.

Sealants are flexible materials used to seal joints between precast concrete units and between such units
and adjacent materials.

Quirk miter is a corner formed by two chamfered members to eliminate sharp corners and ease alignment.

Sealers or protective coatings are clear chemical
compounds applied to the surface of precast concrete
units for the purpose of improving weathering qualities or reducing water absorption.

Retarder, surface is a chemical applied to the mold
surface, used to retard or delay the hardening of the
cement paste on a concrete surface within a time period and to a depth to facilitate removal of this paste after the concrete element is otherwise cured (a method

42

for producing exposed aggregate finish).

ARCHITECTURAL PRECAST CONCRETE

Set-up is the process of preparing molds for casting
including installation of materials (reinforcement and
hardware) prior to the actual placing of concrete.

STATE-OF-THE-ART

1.6 Definitions

Sequential casting see Two-stage precasting.
Shear wall is a wall designed to transfer lateral forces
acting parallel to the face of the wall, from the superstructure to the foundation.
Shop drawings are graphic diagrams of precast concrete units and their connecting hardware, developed
from information in the contract documents. They
show information needed for both field assembly
(erection) and manufacture (production) of the precast
concrete. They are normally divided into:
	Erection drawings—All drawings used to define
the shape, location, connections, joint treatment,
and interfacing with other materials for all precast
concrete units within a given project. Special handling instructions and information for other trades
and the general contractor are also shown.
	Anchor setting or contractor’s setting drawings
—Giving the location of all anchoring hardware cast
into or fastened to the building or structure.
	Production drawings—The actual detail drawings
necessary for production of the precast concrete
units. Such drawings may be set-up drawings or reinforcement and hardware drawings. They should
completely define all finish requirements and include
details of all materials used in the finished precast
concrete units. They are normally not submitted for
approval.
Shrinkage is the volume change in precast concrete
units caused by drying that normally occurs during the
curing and initial life of concrete members.
Side rail is the removable side of a mold.
Strongback is a temporary structural beam or truss
attached to the back of a precast concrete member
to stiffen or reinforce it during shipping and handling
operations.
Structural precast concrete products are units normally produced in standard shapes that carry dead and

1

live load or another units weight. Architectural treatments may be provided on the surfaces of these structural elements, and these should be specially listed in
Contract Documents. Quality assurance for structural
precast and structural precast with an architectural
finish is defined in PCI Manual for Quality Control for
Plants and Production of Structural Precast Concrete
Products (MNL-116).
Thermal movement is the volume change in precast
concrete units caused by temperature variations.
Tolerances are (a) the permitted variation from a basic dimension or quantity, as in the length or width
of a member; (b) the range of variation permitted in
maintaining a basic dimension, as in an alignment
tolerance; and (c) a permitted variation from location
or alignment. There is no intent to split tolerances between structural and architectural tolerances on the
basis of finish or color. Finish and color are separate
issues related to project aesthetic requirements.
Tooling refers to most of the manufacturing and service processes preceding the actual set-up and casting
operations.
Two-stage precasting is the casting of large or steep
returns as separate pieces to achieve matching highquality finishes on all exposed faces and then joining
the separate pieces with dry joints to allow the separate castings to appear and perform as one homogeneous unit.
Unit (member, element) in this document, a precast
concrete piece, or component.
Weatherproofing is the process of protecting all
joints and openings from the penetration of moisture
and wind.
Weather sealing is the process of treating wall areas
for improved weathering properties; see also Sealers.
Wythe is a continuous vertical section of wall tied to its
adjacent vertical element (part of a composite wall).

ARCHITECTURAL PRECAST CONCRETE

|

43

2

Juarez Complex
Mexico, D.F., Mexico;
Architect: Legorreta + Legorreta;
Photo: José Ignacio González Manterola.

CHAPTER TW O
DESIGN CONCEPTS RELATED
TO USAGE AND ECONOMICS
2.1 G
 ENERAL COST FACTORS
The structures reviewed in Section 1.2 illustrate the
wide range of projects utilizing architectural precast
concrete. During the conceptual stage the designer
must consider the cost implications of material selections, textures, surface geometries, cross-sections, piece
sizes, unit repetition, and erection methods. The variations in scope, complexity, and detailing make it difficult to provide accurate cost information for a project
in terms of price per square foot (m2) of wall area prior
to completion of the design concept.
After a design has advanced to the schematic stage,
and the general shapes, colors, and finishes have
been defined, more accurate cost estimates can be
provided. Until this stage is reached, architects are
encouraged to seek the advice of precasters and
consultants. Selected guidelines regarding cost of
architectural precast concrete are included in this chapter for further assistance.
The architect who desires a more detailed understanding of the cost factors involved in precast concrete construction is advised to study both Surface
Aesthetics in Chapter 3 and Design in Chapter 4.
Many of the recommendations in these chapters are, in
the final analysis, based upon considerations of economy. Chapter 6, dealing with Guide Specifications,
will highlight the items that should be included in the
specifications in order to define the optimum quality
for a specific project. This in turn should
help in obtaining accurate proposals from
potential bidders.

interdependent on each other. For example, a cost-efficient sculptured or intricate design may be achieved
within a limited budget by selecting economical concrete mixtures and finishes combined with repetitive
units and efficient production and erection details.
The small college library shown in Fig. 2.1.1 is clad in
exposed aggregate architectural precast concrete. The
smooth horizontal banding and window trim detailing
contrasts with the rough texture of the exposed aggregate facing. The effect achieved is similar to that of
the molded stone banding contrasting with the rough
cut stone on many of the surrounding older campus
buildings. Use of repetition made the precast concrete
panels more cost effective than cut stone.
Repetition in panel design is also the key to achieving
quality and economy in the design of walls. During the
design stage, the exterior walls of a typical office building can be analyzed at three basic locations:
(1) At lower level floors, normally the ground floor
and mezzanine are where significant architectural
expression and detailing will occur;
(2) At all typical floors where repetition of panel size,
shape, and finish occur; and
(3) At the top floor, parapet, mechanical floors,
and penthouse, where there is a likelihood of
increased panel length and often absence of window openings.

The nature of precast concrete is such
that nearly anything that can be drawn,
structurally designed, and readily transported can be constructed. To do this
within reasonable and stated cost limits
requires careful consideration of design
and detailing. Several cost factors influencing architectural precast concrete are
Fig. 2.1.1
Olin Library, Kenyon College, Gambier, Ohio;
Architect: Shepley Bulfinch Richardson and Abbott;
Photo: Nick Wheeler/Wheeler Photographics.

ARCHITECTURAL PRECAST CONCRETE

|

45

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.1 General Cost Factors

A cost breakdown of these three locations will usually illustrate that the lowest wall square foot cost is for
those on the typical floors because of the volume and
repetition. If the unit cost of either location 1 or 3 appears expensive in relation to location 2, a review may
be useful, unless the areas involved are insignificant in
relation to the overall area of the building. The elements
which may be considered in such a review of locations
1 and 3 follows:
Walls for lower level floor(s)
The optimum solution for a ground floor would be a
design that allows the precast concrete units to be cast
in the same master mold as contemplated for the typical floors. This holds true even with the larger openings,

doors, or entrance areas normally required for such a
floor. The illustration in Fig. 2.1.2 is a good example of
this. This convention hotel has a 7-story podium and a
27-story hotel tower. An important design criteria was
to create an image for the building compatible with
the highly textured and detailed finishes on adjacent
landmark buildings. Panel configuration varied from
column/lintel “stick” construction and framed flat
panels at the building podium public spaces to single
and double flat panels with openings for field-installed
windows. Typical window panels are horizontal, double openings at the flat tower face and vertical, double
openings at the curved turrets with false joints in each
panel to create the appearance of stone construction
and uniformity of panel sizes. Panels were configured
to the maximum size that could be transported by
semi-trailer and to the maximum weight that could be
hoisted by crane.
If the ground floor design does not lend itself to the
master mold concept in relation to typical floors, it is often the simplest solution to use precast concrete solely
as cladding consisting of flat sections with returns only
as necessary. One flat mold can be the master mold
and modifications can be obtained by relatively simple
adjustment of bulkheads. Flat panels may also be economical because the layout of a ground floor often
precludes much repetition.
A cladding solution with flat panels and minimum returns may lead to numerous, and fairly small, panels
contradicting later recommendations in this manual.
Smaller flat units in large numbers may still be more
economical than larger units with little repetition
and high tooling costs. Erection of the larger number of units may be partially offset by the use of small
cranes or forklifts, which are often adequate for lower
elevations.
A cladding solution for a ground-floor level design
often allows the choice of more expensive finishes for
this area only. Consideration should be given at the
ground-floor to the proximity of the units to pedestrian traffic. In some cases accumulation of snow or dirt
against ground-floor panels may influence choice of
finish such as natural stone veneer-faced precast concrete, because ease of cleaning becomes an important
factor. The finish requirements for the typical floors
Fig. 2.1.2 Sheraton Chicago Hotel & Towers, Chicago, Illinois;
Architect: Solomon Cordwell Buenz & Associates Inc.;
Photo: Jim Hedrich ©Hedrich Blessing.

46

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.1 General Cost Factors

Mechanical Floors

2

Fig. 2.1.3
Colony Square
Atlanta, Georgia;
Architect: Jova/Daniels/Busby Inc.;
Photo: Jova/Daniels/Busby Inc.

Typical Floors

Mechanical Floors

Typical Floors

Ground/mezzanine
floors

may be less demanding, suggesting a different finish,
provided the combination is aesthetically pleasing or
has a logical separation.
Walls for top floor(s)
The architect should balance the shape and size requirements for the top floor panels with a reasonable

utilization of molds being used on the rest of the project. The top floor is not the place to use excessively
large units unless the design and budget warrant the
additional crane cost. Figure 2.1.3 illustrates the use
of units that are narrower in width than those of the
lower floors.

ARCHITECTURAL PRECAST CONCRETE

|

47

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.1 General Cost Factors

(a)
Fig. 2.1.4(a), (b) & (c)
Crescent VIII Office Building
Denver, Colorado;
Architect: Barber Architecture;
Photos (b&c) LaCasse Photography.

(b)

(c)

The project in Fig. 2.1.4 shows an application of
some of the major points made in this chapter. The
center bay of the building contains stairs, special service areas, and an elevator core (Fig. 2.1.4[a]). The core
was designed to transmit all horizontal loads to the
foundation. The use of integrated, loadbearing architectural precast spandrels and a precast concrete core
facilitated a tight construction schedule. The total precast concrete structure was erected in only eight weeks
with two cranes. The horizontal mass of the building
is broken by expressed vertical pilasters (Fig. 2.1.4[b]),
which break the building into a series of regular bays,
highlighted by granite and concrete accent medallions.
Details such as reveals and medallions also help to reduce the building’s scale and offer visual interest at the
pedestrian level (Fig. 2.1.4[c]). The architect wanted to
maximize the window space while also maintaining a
heavier, substantial wall form. This was achieved with
the detailing and implication of the beam-column look
of the precast concrete panels.

48

|

ARCHITECTURAL PRECAST CONCRETE

The architect’s challenge for the project in Fig. 2.1.5(a)
was to come up with an exterior building skin for a

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.1 General Cost Factors

(b)

Fig. 2.1.5(a), (b), (c)
Shannon Oaks
Cary, North Carolina;
Architect: Cline Davis Architects PA; Photos: Cline Design Associates.

2

the architect and owner to develop a very
detailed, typical selfsupporting wall panel
10 ft (3 m) wide x
34 ft (10.4 m) high.
Great care was taken
to build a mold that
could be used over
and over again, but
would give the project a wonderful sense
of detail and richness
(Fig. 2.1.5[b] and [c]).
Because of the repetitive nature of the panels and their ease of
installation, the precast concrete system
ended up costing less
than brick and steel

(a)

new two-story, 50,000 ft2 (4600 m2) commercial office building that captured the essence of Neo-French
Classical architecture while staying within a fixed market rate cost for this type of structure. The precaster
was brought in at an early design stage to work with

Typical Full Width Panel

Half Width Panel

Typical Corner

Fig. 2.1.5(c) Precast panel elevations and plan details.

ARCHITECTURAL PRECAST CONCRETE

|

49

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.1 General Cost Factors / 2.2.1 Repetition

construction. The result of using this architectural precast concrete system was a beautifully detailed Class
A office building that was 90% leased within four
months.
Coordinated design, complete dimensioning, and clear
specifications (see Chapter 6) are also important factors in
obtaining optimum quality and economy using architectural precast concrete. In the preparation of the contract
documents, the selection and description of materials
and performance requirements should be clearly stated.
They should not be left open to variable interpretations,
however nor should they be overly restrictive.
The contract documents should make reference
to the PCI Manual for Quality Control for Plants and
Production of Architectural Precast Concrete Products
(MNL-117), which includes Category A-1 certification
of the production facility, as the industry guideline for
production of architectural precast concrete elements.
Exceptions to this standard or specific requirements
should be clearly set forth in the contract documents.

2.2 D
 ESIGN ECONOMY
Understanding architectural precast costs is essential to designing affordable façades that enhance the
overall building design while meeting the owner’s budget. Understanding the architectural precast concrete
manufacturing process can help achieve design goals
and control costs.
During a project’s conceptual stage, the designer has
many variables to consider that affect precast concrete
cost. A local precaster can assist with preliminary design and budget estimating early in the project’s design
phase. Piece size and repetition typically have the most
significant cost impacts. In addition, material selection,
textures, surface geometry, cross-section, erection details, jobsite access conditions, and connections can
affect cost. The custom, sculptured designs that are
possible with precast concrete may be achieved within
a limited budget by selecting appropriate aggregates
and textures combined with repetitive units and efficient production and erection details. Input from the
precaster can be beneficial in developing options for
creating an economical design that also satisfies the
designer’s aesthetic requirements.
During preliminary design, a precast concrete project
can be preliminarly budgeted on a square-foot (m2) basis. Although this provides a starting point, it is recom-

50

|

ARCHITECTURAL PRECAST CONCRETE

mended that the designer seek additional estimating
assistance from a precaster. Working with a precaster
on a specific project will help determine a final budget
that is more accurate than a ballpark price per square
foot (m2). A cost per square foot (m2) can be misleading to general contractors and architects because
square foot (m2) quantities are calculated differently
from precaster to precaster depending on the take-off
procedures. Also, total work scope requirements such
as site restrictions, work scope inclusions, and detail
manufacturing requirements are initially unknown.
Budget pricing from local precasters, submitted in
writing and including assumptions, will aid design efforts from schematic design through final contract documents. As a project evolves from preliminary sketches
through working drawings, the precaster(s) should be
informed of all changes.
Pricing accuracy depends on the information provided
to the precaster’s estimator. This discussion on design
economy uses square foot (m2) prices to describe a designer’s precast concrete options. All prices are for relative comparison only and should not be used to make
decisions for individual projects.
The design and detailing of the precast concrete units
should reflect good production concepts. Consultation
with a precaster at an early stage will be helpful. The
designer needs to define the shape of the units and
their appearance.

2.2.1 Repetition
A key element to cost-effective production is minimizing the number of molds and mold changes, and maximizing the number of castings from each mold, particularly if the molds have shape. Efficiency and economy
are achieved by making it possible for similar, if not
identical, shapes to be produced from the same basic
(master) mold, and by minimizing the time required to
disassemble a mold and reassemble it for the manufacture of the next piece. Figure 2.2.1(a) shows the master
mold for the production of the arch member panels for
the project in Fig. 3.3.18(a), page 121. The largest segment of the arch is shown in Fig. 2.2.1(b).
Careful planning is necessary to achieve good repetition in the design without sacrificing design freedom.
For example, many design variations may be developed by incorporating two basic architectural panel
types (spandrel panels and floor-to-floor panels with

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.2.1 Repetition

openings) on the same structure. These panel types
may also be varied with different architectural finishes
and textures. Attention should be focused on the overall geometry of the structure, not only on the shape
of the panel. The cost of complex shapes becomes
economical through repetitive precasting as the invest-

2

ment made in fabricating a complex mold is amortized
over a greater number of pieces. Occasionally, due to
production schedule compliance, a precaster may need
to construct multiple molds to produce the required
number of panels within a certain time period.

Fig. 2.2.1(a) & (b)
The master mold for the production of the arch member panels
for Fig. 3.3.18(a) page 121.
Jefferson Pilot
Greensboro, North Carolina; Architect: Smallwood, Reynolds,
Stewart, Stewart & Associates; Photos: Dean Gwin.

(a)

(b)

ARCHITECTURAL PRECAST CONCRETE

|

51

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.1 Repetition

(a)

pecially on the tower portion of the building. This resulted in more-pronounced shadow lines, providing
enhanced visual definition on the tower portion of
the building. A strong pilaster expression, enhancing
the vertical planes of the building, is also achieved
through the use of precast concrete. Combined with
the deeply recessed windows, the articulation of the
pilaster forms provides a dimensional texture to the
entire building.
It is often the case that, in the initial design stage,
a high degree of repetition appears possible.
However, as the details are finalized, considerable
discipline is required on the part of the designer if
the creation of a large number of non-repetitive
units is to be avoided. Budget costs used at the initial design stage should take into account the possibility that the number of different units will increase
as the design progresses. If non-repetitive units are
unavoidable, costs can be minimized if the units can
be cast from a “master mold” with simple modifications without the need for completely different
molds. However, even relatively minor variations,
such as a dimensional change of a rail, blockout location, connection hardware position, or a different
number of blockouts of any kind, are mold changes
that increase costs.

(b)
Fig. 2.2.2(a) & (b) Trinity Place, Boston, Massachusetts;
Architect: CBT/Childs Bertman Tseckares, Inc.; Photos: Edward Jacoby.

The multifamily residential condominium with street-level
retail space is situated at the opening of a major thoroughfare between the two distinct city districts (Fig. 2.2.2[a]).
The building’s precast concrete façade was chosen as a
means of integrating this building with the surrounding
neighborhood. The ability to control the color and texture
of the finish, and the ability to break up the façade into
smaller elements with rustication joints, allows the precast
concrete to relate comfortably to both the 19th- and 21stcentury buildings that surround it. The use of precast concrete provided the ability to create a prefabricated window
anchor system throughout the building, which enabled the
creation of multiple visual elements. On the lower floors,
window boxes protrude from the façade, bringing the
building to life for pedestrians (Fig. 2.2.2[b]). Elsewhere, the
designers were able to achieve deep window recesses, es-

52

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.2.1 Repetition

Fig. 2.2.3(a) & (b)
111 Huntington Avenue/Belvedere Residential
Boston, Massachusetts; Architect:CBT/Childs Bertman Tseckares,
Inc.; Photos: Jonathan Hillyer.

2

(a)

(b)

The project in Fig. 2.2.3(a) is an integral part of a substantial redevelopment of an outdated 1960s urban complex. The major elements include a 36-story office tower;
an 11-story, 61-unit residential building; and an enclosed
winter garden pedestrian arcade. The residential building
follows the curving street edge, taking a graceful crescent
shape. The precast concrete lends itself to the fluidity of
such a design, as the convex base of the building gently
turns inward to form a concave surface. The consistency
of the precast concrete was a necessity in the construction of the circular design, which requires repetition in
form to achieve its desired look (Fig. 2.2.3[b]). The precast
concrete takes on a columnar form around the entrance
to the winter garden, and continues along the garden’s
entire 480 ft (146 m) length, providing a rhythmic pattern
both internally and externally on the adjacent park. The
precast concrete base to the office tower skillfully blends
with the steel and glass construction. The base integrates
vertically with the tower that soars above it, as precast
concrete “fingers” reach upward into the highrise portion
of the structure as the material transitions. The precast
concrete forms a delicate frame for the window walls and
incorporates scale, dimension, and shadow to the wall.

If unfamiliar with architectural precast concrete, prior to
designing wall panels, the architect should visit an architectural precast concrete manufacturing plant, as well as
any projects that are under way. This way the designer can
become familiar with the manufacturing processes and
installation procedures and, most importantly, establish
realistic expectations for the finished product. Elements
such as the fabrication of molds, challenges to casting and
finishing specific designs or shapes, relative material costs,
handling methods at the plant and jobsite, approaches
for connecting panels to a structure, and establishing acceptable color ranges are important to fully understand
precast concrete and maximize its potential.
Reveals and rustications must be placed in a repetitive
pattern in order to minimize modification throughout a
mold’s life. Reveals, like all form features, must be designed
with draft (by creating bevels) so the panel can be stripped
from the mold without damaging the mold feature.
Cost premiums are introduced to a project when the
panel cross-section becomes more complex or intri-

ARCHITECTURAL PRECAST CONCRETE

|

53

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.1 Repetition / 2.2.2 Mold Costs

The point behind designing repetitive pieces is to amortize engineering and mold costs effectively. As many
pieces as possible should be designed to be cast in the
same mold and produced from a single shop drawing.

Removable bulkheads

Master mold

Panel with left
hand return

Panel with right
hand return

2.2.2 Mold Costs

Intermediate panel

The architect can make a significant contribution to economic production by designing precast concrete panels
with knowledge of the master mold concept and by providing the precaster with sufficient production lead-time,
making the duplication of molds to meet project schedule
requirements unnecessary. The concept is simply to design
the largest possible mold for a particular unit, whereby
several variations from the same mold can be produced
by varying mold component accessories. Units cast in this
mold need not be identical, provided the changes in the
units can be accomplished through pre-engineered mold
modifications. These modifications should be achieved
with a minimum change-over time and without jeopardizing the usefulness or quality of the original mold. Typical
applications are shown in Fig. 2.2.4. When using a master
mold, individual castings do not have to be the same color
or texture.
It is relatively easy to alter the panel size if the variations can be contained within the total master mold
envelope. This strategy eliminates the need (and cost)
of constructing a mold for every panel change. The
use of bulkheads, blockouts or reveals placed on top
of the mold surface is less expensive than cutting into
the mold surface for a projecting detail.
When a large number of precast concrete units can be
produced in each mold, the cost per square foot will be
more economical (Table 2.2.1 [mold cost is for illustration
only]). The master mold concept is illustrated in Fig. 2.2.4.

54

|

ARCHITECTURAL PRECAST CONCRETE

First floor panel

Mold cost depends on size, complexity, and materials used. The mold material selected and number of
molds depend on a project’s schedule. A project with a
long precast concrete production period should permit
fewer molds to be built.

Typical floor panel

cate surface features are added. Projecting cornices,
bullnoses, form liners, bottom and/or top returns, and
curves are the most typical features to be added. The
exact sizes, shapes, and locations are the designer’s options. Cost will be added if the location of these features
within a mold will be changed frequently. On the other
hand, these intricate features can be added at a minimal
cost if they are used repetitively in the overall design.

Top floor + parapet panel

2

Typical panel
Corner panel

Fig. 2.2.4 Views of master mold.

Number of
uses

Panel size,
ft2

Mold
cost

Mold Cost
per ft2

1

200

$5000

$25.00

10

200

$5000

$2.50

20

200

$5000

$1.25

30

200

$5000

$0.83

Table 2.2.1. Effect of repetition on panel mold square foot cost.
(Mold cost is for illustration only.)

In these examples, a large number of panels can be produced from a single mold, built to accommodate the largest piece and then subdivided as needed to produce the
other required sizes. Whenever possible, the largest pieces
should be produced first to avoid casting on areas that have
become worn and damaged by placing and fastening side-

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.2 Mold Costs

form bulkheads. Although every project will have some
atypical conditions, the most successful and cost-effective
projects maximize the repetition of elements. The more often a mold is re-used, the lower the unit cost of the piece.
This means that careful planning is necessary to achieve
good repetition without sacrificing design freedom.
The premium cost for complex shapes can be controlled by adding details to specific forms only. Examples
include designing a cornice at parapet panels, a sill detail at intermediate floors, or one elevation as a radius.
An example of pre-engineered mold changes is shown in
the models for a loadbearing panel in Fig. 2.2.5. The forms
were made in two pieces that bolt together. The head section was removable and
could be replaced with a
modified upper fascia to
achieve design variations.
The process is illustrated
starting with the two basic
panels joined at the jobsite
with the addition of a corner unit (Fig. 2.2.5[a]). By
redesigning the upper fascia only, the panel takes on
(a)
a new look, while keeping
costs substantially lower,
rather than if a complete
new mold were needed
(Fig. 2.2.5[b]). All of the
engineering of the loadbearing section remains
unchanged. Even though
the panel in (a) is essentially
the same as (b), the slight
change in height and the
(b)
deletion of fins creates a
different appearance. Fig.
(c)
2.2.5(c) shows the same
panels with no fins and
varying fascia heights.

Fig. 2.2.5 Pre-engineered mold
changes.

The wide flat center section of the panel can be
blocked out for full windows, window frames, or
doorways. Also, different
types of concrete finishes
and textures are possible in the center area.

2

A demarcation groove makes casting of the different
finishes a simple operation. By having the flexibility of
varying expressed material usage, it is possible to relate
the basic tone of the design to the existing surrounding structures.
Present bidding practices often result in late award of
precast concrete subcontracts. The number of molds of
a given type required for a project often is determined
by time allowed for completing the project. In many
cases, this time factor to meet the project’s schedule is
what creates the demand for duplicate molds, overriding the desire for mold economy. The necessity for extra
molds increases costs and partially offsets the intent of
designing for high repetition. The designer should discuss realistic precast concrete engineering and production lead times for the project with a precaster. It is vital
to promptly respond to requests for information (RFIs) to
ensure that shop drawings or, at least, shape drawings
are approved and mold manufacture begins on schedule. Precast concrete production lead time is an activity that cannot be effectively compressed. Overrunning
lead time will impair production, delivery, and, hence,
construction.
It is vital to include precast concrete scheduling information with the bid documents. This will ensure all
bidders understand the project time frames that are
required. Ample lead time also will allow the manufacture of larger pieces first, followed by smaller ones,
thus minimizing the cost of form repairs.
If the precaster is not provided with the full lead time
given to the general contractor, the cost of the precast
concrete job invariably rises or the project schedule
must often be lengthened. It is desirable to specify that
major subcontracts, including precast concrete, be let
within a short, defined time period after the general
contract is awarded.
Pre-bidding and awarding contracts authorizing engineering and drafting costs (subject to stated project
cancellation fees) can result in both production schedule time savings and savings in production costs and
are recommended. Part of this savings can be passed
on to the owner directly. Pre-awarding precast concrete
subcontracts also benefits the owner by decreasing the
start-up time required after award of the general contract. Such pre-awarded subcontracts may then be assigned to the successful general contractor.

ARCHITECTURAL PRECAST CONCRETE

|

55

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.3 Other Forming Considerations

2.2.3 O
 ther Forming Considerations
In addition to considering maximum form reuse, the
final design should take into account ease of removal
from forms. This allows the precaster to efficiently meet
schedules and budgets without impacting the design
aesthetics.
Optimum economy in production is attained if the
panel can be separated from the mold without disassembling the mold. This is done by providing draft
(slope) on the sides of all openings and edges and by
eliminating or minimizing forming the panel’s back
face (upside in the form). Drafts are a function both of
shape and production techniques (see Section 3.3.2).

Fig. 2.2.7(b)

Fig. 2.2.7(c)

Molds. The complete envelope mold is a box mold
where all sides remain in place during the entire casting
and stripping cycle. Figure 2.2.6 shows a generic envelope mold designed with the minimum workable drafts. A
complete envelope mold is shown in Fig. 2.2.7(a) and the
corresponding panel in Fig. 2.2.7(b). The quality of joints
produced from such molds is shown in Fig. 2.2.7(c). Figure
2.2.8 illustrates a mold without edge draft requiring removable side and end bulkheads for stripping.

Ribs, reveals
or false joints

Side
forms

12

6

Panel projections

8

Fig. 2.2.6 Total envelope mold—minimum positive draft.

Fig. 2.2.7(a)
Lift panel

Lift panel
Remove
bulkhead

Flat panel

Fig. 2.2.8 Conventional molds.

56

|

ARCHITECTURAL PRECAST CONCRETE

Remove
bulkhead

Sculptured panel

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.2.3 Other Forming Considerations

The configuration of the envelope mold requires that
panels be stripped flat from the mold and rotated to a
vertical position later. The slightly wider joint between
panels caused by the draft required on the side forms
is a potential disadvantage of envelope molds.
Several modified versions of the complete envelope
mold that will accommodate precast concrete units without drafts along one or more edges are illustrated in Fig.
2.2.9. Because the loose side rails or back forming are
stripped with the unit, the mold allows 90 degree returns
or returns with negative drafts to be cast using an envelope mold. The modified envelope molds are much easier
to reassemble than loose bulkhead forms, because daily
measuring and aligning of the side rails is not necessary.
When properly designed, a modified envelope mold will
provide the same good corner details and high-quality finish found on units cast from a complete envelope mold.
Thought must be given to preventing leakage of fresh
concrete paste especially where removable side or end
rails attach to forms. A point of leakage can mar the finish. A return as indicated in Fig. 2.2.9(c) and (d) will cause
any potential leakage to take place where it will not be
seen in the finished product.

2

Architectural precast concrete units normally are cast
in a horizontal or flat position with the exposed, textured, or sculptured face down to give the maximum
aggregate consolidation at the panel surface. Twosided precast concrete pieces (front and back) requiring identical appearances should be avoided, as facing
aggregate will need to be seeded on the top surface
resulting in likely texture variations.
Where the shape requires it, the form may be made
in parts with removable sections (such as side rails and
top forms) that must be assembled and disassembled
with each day’s concrete placement (Fig. 2.2.10[a] and
[b]). This has the effect of increasing panel cost.
Fig. 2.2.10(a) An envelope mold with haunches on back.

Wood side rail

Adjustable clamps

Wood mold face

Nail

Casting
deck
Wood block
Wood end gate
(a) Spandrel panel

(b) Column cover

Wood side rail
Wood wedges
Wood wedge rail

Fig. 2.2.10(b) Removable sections within a mold.

Lift panel

Examples of face and back forming
for deeply sculptured panels
Lift panel
Removable bulkhead

2

1

Removable side rail
2

1

3
/4"
min.

3
/4"
min.

/8"

1

/8"
(c) Flat panel
1

Fig. 2.2.9 Modified-envelope molds.

(d) Sculptured panel

Back forming is used to create returns that give the
appearance of thick, massive panels that add significant shadow features to the façade. These returns also
can allow windows to be set back away from the building’s face. To achieve these shapes, special forms must
be constructed and then suspended over the primary
mold to create the desired panel depth.
A common production method to make large returns is a
two-stage concrete placement (see Section 3.3.9). The return piece is produced on Production Day 1. On Day 2, the

ARCHITECTURAL PRECAST CONCRETE

|

57

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.3 Other Forming Considerations / 2.2.4 Panel Size and Panelization

return piece is removed from its form and is connected to a
master mold. The return is cast monolithically to the master piece. Two-stage concrete placements create a more
uniform texture on all sides of the panel. A quirk should
be provided in the corner so the return is not formed to a
sharp edge that is easily chipped.
The details for casting individual panels should always
be left to the precaster. Elevations, wall sections, and
details of each different type of wall panel should be
drawn by the architect. When using large elements,
if the appearance of smaller panels is desired for aesthetic reasons, false joints (rustications) can be used to
achieve this effect.
At times a compromise may be required between
the finish and the shape of a precast concrete panel.
Wherever possible, the designer should avoid fragile
edge details. Chamfered or eased edges reduce edge
damage and mask minor irregularities in alignment.

2.2.4 P
 anel Size and Panelization
Precast concrete pricing is determined primarily by
the size of the pieces and repetition. Pricing is more
dependent on large pieces than on a large project. For
example, a 100-piece project of large panels can be
less expensive per square foot (m2) than a 1000-piece
project using much smaller panels.
The reason piece size is so important is because most
labor functions performed by an architectural precaster
and erector are required because of the existence of a
piece. The more pieces the project has, the more labor
hours it will take to engineer, cast, strip, finish, load,
deliver, and install the panels. Therefore, it is more economical to enclose a larger portion of the building’s
exterior with fewer precast concrete panels (see also
Section 3.3.10).
For maximum economy, minimize the number of
pieces by making them as large as possible within normal manufacturing and shipping limitations. Handling
and erecting precast concrete components constitutes
a significant portion of the total precast concrete expense. The cost difference in handling and erecting a
large rather than a small unit is insignificant compared
to the increased square footage of a large unit, Table
2.2.2. To be economical, a project’s average piece size
should be at least 100 to 150 ft2 (9 to 14 m2) and, ideally, larger than that.

58

|

ARCHITECTURAL PRECAST CONCRETE

There is no exact optimum panel size. Usually the
optimum panel size is dictated by size and weight
limitations imposed by transport (for example, weight
restrictions and bridge or power line clearances), site
access, or crane capacity. The panel size is also a function of the design loads and support locations for connections. Close collaboration between the designer
and a precaster is required during the early stages of a
building’s design to determine the optimum panel size
or panelization scheme. Piece sizes that require highway permits for over height, width, length, or weight
generally should be avoided.
There is a balance between maximizing potential
economy of the façade elements and maintaining the
economy of the supporting structural system. The key
is to recognize where localized loads will occur. Often
the added cost of local reinforcing of the supporting
structure that may be required to accommodate larger
precast concrete panels will be more than offset by
savings that result from erecting fewer panels.
The designer can ensure a good average piece size by
spanning a full bay with spandrels, and designing multistory column covers and large wall panels. Designing
larger panels, even though they may carry a hauling
premium, may be the most cost efficient. For example,
an office building with 30 x 30 ft (9.1 x 9.1 m) column
spacing requires fewer columns and concrete panels
and yields a more wide-open interior than the same
building with a 20 or 25 ft (6.1 or 7.6 m) column spacing. The cost premium (if any) to haul two 30-ft-long
(9.1 m) panels versus three 20-ft-long (6.1 m) panels
usually can be more than overcome by cost savings in
other manufacturing areas like engineering, production, and installation. The typical parking structure
may have perimeter panels that are 60 ft (18.3 m) long
Panel size,
ft2

Erection cost per piece,
dollar amount per ft2
$500

$1000

$1500

$2000

50

10.00

20.00

30.00

40.00

100

5.00

10.00

15.00

20.00

150

3.33

6.67

10.00

13.33

200

2.50

5.00

7.50

10.00

250

2.00

4.00

6.00

8.00

300

1.67

3.33

5.00

6.67

Table 2.2.2. Effect of panel size on erection cost per square foot.
(Erection costs are for illustration only.)

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.2.4 Panel Size and Panelization / 2.2.5 Material and Labor Costs and Uniformity of Appearance

or more running parallel to the 60-ft-long double-tee
floor system. The double tees cannot carry the perimeter panel weight. These panels must be designed to
span column to column. Therefore, the added cost to
haul a 60-ft-long panel is offset by the omission of a
supporting structural beam.
In addition to providing cost savings during erection,
larger panels provide secondary benefits by shortening
a project’s schedule, reducing the joints to be sealed,
and requiring fewer connections. Using fewer lineal
feet of joint sealant translates directly into less probability of joint sealant failure, which could allow both
air and moisture infiltration into the wall cavity. Thus,
large units are usuably preferable unless they lack adequate repetition or incur significant cost premiums for
transporting and erecting.
If design of an elevation requires the appearance of
smaller units that are not economic, the inclusion of
false joints (rustications or reveals) cast into the face
of larger elements can give the illusion of smaller elements and provide a solution that improves overall
economy. These false joints can be caulked to match
the appearance of actual panel joints to increase the
illusion of small panels.

2.2.5 M
 aterial and Labor Costs and
Uniformity of Appearance
Table 2.2.3 lists material factors and labor processes
that affect both cost and the uniformity of appearance.
It is difficult to provide representative cost figures for
different precast concrete surface finishes, because individual plants may price them somewhat differently.
Some plants, for instance, consider an acid-etched
surface an expensive finish, particularly if they infrequently use this method of finishing. Some precasters
discourage its use, while others may prefer its use to
sandblasted or retarded (exposed aggregate) finishes
to obtain a similar appearance.
The cost of cement in the finished, erected product
will normally vary between 3 and 6% of the total cost
per square foot of wall, depending upon the concrete
volume per square foot of wall and whether gray or
white cement is required. The premium for white cement is not a great percentage of the overall cost,
although white cement is from 2 to 2.5 times more
expensive than gray. In addition to specifying white cement for color effect, the architect will also obtain bet-

2

ter uniformity in color than is possible with gray (see
Section 3.2.1).
Pigments have a small impact on the overall cost of
a project. For an architectural precast concrete panel
with a 3 in. (75 mm) face mix, pigments will generally
add $0.10 or less per square foot for most light, earthtone colors and $0.40 or less for intense or dark colors
such as charcoal, chocolate, bright red, or orange. The
exceptions are green and blue, which need to be investigated on an individual basis, because their cost may
impact a project substantially.
Aggregate selection. For reasons of appearance and
cost, aggregate choice is an important factor. Aggregate
cost is determined primarily by the transportation charges from the quarry to the precast concrete plant. Most
aggregates cost about the same to remove from the
earth and to crush to the appropriate size. The trucking cost from the quarry to the plant is the principal
cost variable. In order to minimize mix cost, a designer
should discuss aesthetic requirements such as aggregate
color options and their associated costs with a local precaster. Local aggregates, such as gravels, should not be
overlooked. They will be more economical and may look
very attractive with the proper matrix and finish. The
cost of aggregates should be presented on the basis of
both per cubic yard (m3) and per square foot (m2). Even
the most expensive aggregates are often practical in
exposed aggregate concrete, especially when they are
used only in the thin face mix. For example, a 2 in.(50
mm) thick layer of face mix will use approximately 15
lb (6.8 kg) of coarse aggregate per square foot. Cost of
aggregate will be approximately $0.20 per square foot
for each $25 per ton of delivered aggregate cost to the
precast concrete plant.
A particular aggregate’s cost should be calculated
only for the amount of face mix used. If a gray backup mix is used, do not calculate this material cost for
pricing comparisons. Most precast concrete panels are
produced with face mix thicknesses of 2 to 3 in. (50
to 75 mm) and a gray backup mix. Panels with large
projections and returns will increase the face mix quantity required. Window setbacks may dictate the thickness of the face mix. As the set back increases, so does
the amount of the more costly face mix. If the panel
configuration is such that little or no backup concrete
can be used, then the cost of the facing aggregate can
have a significant effect on the cost of the panel.
Because the selection of aggregates (and, to a smaller
degree, that of cement) has a substantial influence on

ARCHITECTURAL PRECAST CONCRETE

|

59

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.5 Material and Labor Costs and Uniformity of Appearance

Table 2.2.3. Factors related to cost and perceived color uniformity.

Assumptions

Comments
Costs will generally decrease as panel sizes increase. The most cost effective panels are
generally larger than 100 to 150 ft2 (9 to 14 m2).

General

Flat panels without recessed windows can use less face mix. Panels with recessed windows,
shape, or returns require more face mix.
Form costs should be amortized over the number of castings that are made within the form.
Always attempt to use the master mold concept.

Cement
content

Aggregates
Pigment
dosage

Gray cement is sold for structural applications. The cement manufacturers do not attempt
to control the color of gray cements. They do actively try to control white cement color and
brightness.
Gray and white cements can be blended to achieve reasonable color uniformity at lower cost.
Color uniformity normally increases as the percentage of white cement increases.
More- and less-expensive local aggregates sometimes are blended to reduce costs.
Lower dosages of pigments often are used to create subtle shades, but very low dosages will
not yield good color consistency. High dosages are used to create strong colors. Often white
cement must also be used to increase pigment effectiveness and, thus, color consistency.
1. Often partial bands of form liners are used to create texture differences within a panel. This
is generally less expensive than covering the entire form surface with a liner.

Form surface

Remember that a form liner is a manufactured product and often has size and module
limitations that must be considered during design.
2. Smaller, less complex projections will greatly reduce the cost.
1. Acid-etched or lightly sandblasted surfaces should normally be created by using a white
cement base for color uniformity.

Surface finish

2. Because of the fine, flat surface resulting from an acid-etched or light sandblast finish, the
panel surface should be broken up with details or rustications to break up the surface mass.
Doing so will result in a more uniform color appearance.

Mix/Finish
complexity

Details must be provided at mix/finish changes within a panel in order to provide a termination
line for a mix/finish change.

final cost, the architect may obtain estimates or bids
with a base price corresponding to the lowest cost
combination suitable for the project, and with quoted
premiums for upgrading of materials. This should not,
however, include more than one or two alternatives,
as the interaction between the costs of the materials,
finishes, shape, and production techniques may complicate evaluation of such premiums.
Precasters can modify concrete ingredients, depending
on the selected finish, in order to lower material costs.
For example, most acid-etched finishes will not expose
the coarse aggregate. Thus, the cost of special coarse
aggregates can be minimized or eliminated since they

60

|

ARCHITECTURAL PRECAST CONCRETE

will not be seen. Since sandblasting dulls the coarse aggregates, less expensive aggregates may be selected.
Local, less expensive aggregates may look very similar to
expensive special aggregates after they are sandblasted.
A bushhammered finish will give a similar appearance to
sandblasting without dulling the aggregates. Exposed aggregate or retarded finishes tend to be more expensive
because they require colorful, premium cost coarse aggregates. This procedure exposes the coarse aggregate
and reveals its natural beauty.
By incorporating demarcation features, multiple mixtures
can be incorporated in a single panel. A designer can also
achieve different colors and textures from a single precast

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.2.5 Material and Labor Costs and Uniformity of Appearance / 2.2.6 Design Options

concrete mixture simply by varying the finish treatment.
This multiple-finish technique offers an economical yet
effective way to heighten aesthetic interest.
Mix Design. It is desirable to develop the mixture
proportions before the project goes out for final pricing. Most precasters are eager to assist the architect in
developing a design reference sample as early as possible. The best method in selecting a sample is to visit the
precast concrete plant to view a multitude of samples
and finished panels stored in the yard. Alternatively, a
designer can refer the precaster to a selection from the
PCI Architectural Precast Concrete – Color and Texture
Selection Guide, to an existing project, or provide a
piece of natural stone (or other material) to match or
refer to.
Often a required panel finish will require a new, one-ofa-kind concrete mix. When visiting a plant, the designer
can select the cement color, aggregate type and size, and
surface finish method/depth. Asking a precaster to make
several different samples for a project is common and encouraged. Once a project’s 12 x 12 in. (300 x 300 mm)
sample for each color and texture has been finalized, the
designer should make the sample available to all interested precast concrete bidders to view and photograph.
In some cases, multiple samples are made so that each
precast concrete bidder can have a sample. Listing the
exact concrete ingredients in the specification is not necessary but encouraged.
Reinforcement and Connection Hardware. The
cost of reinforcement is typically not a significant variable in architectural precast concrete. The amount of
reinforcement is mainly determined by load requirements, such as handling, loadbearing, or other structural functions. An exception is the choice of finish
of the reinforcement and connection hardware. The
cost of galvanized or epoxy-coated reinforcement is
substantial, and is not normally required (see Section
4.4.7). Additionally, it is not a substitute for adequate
concrete cover or concrete quality.
Connection hardware cost is governed mainly by
structural load requirements (including special structural functions, possible earthquake considerations and
weather exposure) and the building’s structural system.
Hardware costs may be minimized by making the precast concrete units as large as is consistent with the size
limitations discussed in Sections 3.3.10 and 4.5.3.
On structural steel buildings, preweld connection

2

hardware can either be attached to the structure in the
field or in the structural steel fabricator’s facility using
precasters drawings. Structural steel bracing to resist
torsion of the structural frame members should be provided and installed by the structural steel fabricator.
Four lateral and two gravity connections are the minimum required for most precast concrete units, regardless of panel size. The labor cost of producing and
handling small individual pieces of hardware normally
exceeds the material costs, thus increasing the relative
cost of hardware for small units.
On steel frame structures, gravity and lateral support
brackets (for precast concrete connections) should be
in the structural steel fabricator’s scope of work and
should be shop welded to the structural steel columns
using precaster’s drawings rather than field welded. It
is much less expensive to shop fabricate and shop weld
them than to hoist and field weld heavy support brackets. Also, the structural steel bracing to resist torsion of
the structural frame members should be provided and
installed by the structural steel fabricator.

2.2.6 Design Options
Design options for precast concrete panels are literally
endless. Employing these options intelligently adds a great
deal of design interest to a project with only minimal cost
increases. The following design strategies can cost from
pennies per square foot to a few dollars per square foot.
1. Incorporate multiple colors throughout a building
façade.
a. Panels can contain more than one concrete face
mix.
b. Panels can be produced with multiple finishes.
The combination of finish methods will determine the cost impact.
2. Add a special shape to one distinct building area.
a. Design an appendage to an existing form. Doing
so will cost less than adding a full form, yet will
create a unique building detail.
b. Set windows back from the building’s face at one
or two column bays, or at certain levels.
c. Add a few small ornate pieces at the entrance or as
site walls. The small panels will be more expensive
per square foot, but a few of them amortized over
the entire project will add minimal additional cost.

ARCHITECTURAL PRECAST CONCRETE

|

61

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.2.6 Design Options / 2.4 Precast Concrete Panels Used As Cladding-General

3. Incorporate brick, tile, terra cotta, or natural stone
accents into the precast concrete.

or the building’s structural grid design. Typical wall panel
system cross-sections are shown in Fig. 2.4.1.

In most cases, design interest can be enhanced without increasing cost by using more complex precast concrete pieces in one area and offsetting the cost premium
by economizing in another area.

The use of precast concrete cladding is a practical and
economical way to provide the desired architectural expression, special shapes, and uniform finishes. When
used over steel columns and beams, cladding can provide the required fire-resistance rating without resorting
to further protection of the steel, under certain conditions. When used over cast-in-place concrete columns
and beams, it will often permit the achievement of a
uniformity of finish in combination with a special architectural shape, all in the most economical manner.
Cladding can be multifunctional, for example, by providing space behind for services and exterior grooves or
buttons for vertical window-washing machinery.

2.3 T OTAL WALL ANALYSIS
The total cost of an architectural precast concrete wall
may be lowered by taking full advantage of the precast
concrete portion. In addition to acting as exterior walls,
the precast concrete panels may perform other functions:
they may be loadbearing, wall-supporting, serve as formwork or shear walls or be used as grade beams; they may
be insulated or may provide the interior finish; they may
serve partly or fully as containers of mechanical/electrical
services; or they may combine several of these functions
to become a wall sub-system. Precast concrete panels
may also be cast compositely with other materials to provide an entirely different finished surface. Clay products
(brick, tile, and terra cotta) and natural stones (granite,
marble, limestone, and sandstone) have all been used as
veneer facing.

2.4 P
 RECAST CONCRETE PANELS USED
AS CLADDING
2.4.1 G
 eneral
The use of non-loadbearing precast concrete cladding
has been the most common application of architectural
precast concrete. Cladding panels are those precast concrete elements that resist and transfer negligible load
from other elements of the structure. Generally, they are
normally used only to enclose space and are designed
to resist wind, seismic forces generated from their selfweight, and forces required to transfer the weight of the
panel to the support structure. Cladding units include
solid wall panels, window wall units, spandrels, mullions,
and column covers. Their largest dimension may be vertical or horizontal. These units generally may be removed
from the wall individually without affecting the stability
of other units or the structure itself. Precast concrete cladding panels can be made in a wide range of shapes and
sizes. For the purpose of discussion, cladding wall units
do not extend in height beyond a typical floor-to-floor
dimension, so floor levels can be defined by horizontal
joints. They are normally limited in width to less than or
equal to the bay width of the structure. The width of the
panel is usually dictated by architectural considerations

62

|

ARCHITECTURAL PRECAST CONCRETE

The extent of repetition and the choice of sizes, shapes,
and finishes are the major design and cost considerations
for cladding units. Panel size and weight for transporting and crane capacity constitute the major dimensional
(panelization) criteria. Economy in the use of precast
concrete cladding is achieved by paying close attention
Fig. 2.4.1 Typical wall systems.
“Z” furring or metal framing to
create 1-in. minimum air space (not
required if no gypsum wallboard)

(a)
Precast
concrete
panel

5
/8-in. thick gypsum wallboard, painted
(optional)

Extruded or expanded polystyrene
insulation, continuous
Precast concrete
sandwich panel
(b)
Precast
concrete
panel

“Z” furring or metal framing to
create 1-in. minimum air space
(optional)
5
/8-in. thick gypsum wallboard,
painted

Extruded polystyrene insulation,
continuous, and taped at joints

Precast concrete with
rigid insulation
(c)
Precast
concrete
panel

1-in. minimum air space
5
/8-in. thick gypsum wallboard,
painted

Batt insulation with kraft paper on
inside face, in zones 3, 4 and 5
(see page 419), in metal framing

Precast concrete with batt insulation

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.4 Precast Concrete Panels Used As Cladding General

2

originated for the requirement of loadbearing walls,
that an area must be provided between glazing to carry vertical loads, and so windows were relatively small.
The reappearance of this pattern derives some rationale from the needs of energy conservation that mitigates against large areas of poorly insulated glazing.
Although today, more energy-efficient fenestration
products allow more freedom in matching daylighting
requirements with the inherent energy efficiency of the
precast concrete units. A much stronger impetus comes
from the dictates of architectural fashion and the desire to return to molded façades and the visual interest
that can be obtained by the traditional manipulation of

Fig. 2.4.2
Las Olas City Centre, Fort Lauderdale, Florida;
Architect: Cooper Carry Inc.; Photo; Ed Zealy.

to the design and detailing of the precast concrete units.
This is a basic requirement of all precast concrete, but
particularly so for units that function only as cladding.
In high-rise buildings, three characteristic façade patterns can be identified that considerably impact the
panel design. The first is cladding that plates the structural framing, vertically and horizontally; the large opening then being infilled with glass (Fig. 2.4.2).
The second pattern eliminates the column covers,
and the façade then becomes alternating horizontal
bands of spandrel panels and glazing (Fig. 2.4.3). In
this pattern, the panels and glazing are placed in front
of the columns.
The third pattern is a return to the traditional façade
design of rectangular window openings “punched”
into a plane surface (Fig. 2.4.4[a] and [b]). This pattern

Fig. 2.4.3
Nashville City Center
Nashville, Tennessee;
Architect: The Stubbins Associates Inc.;
Photo: Jonathan Hillyer.

ARCHITECTURAL PRECAST CONCRETE

|

63

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4 Precast Concrete Panels Used As Cladding General / 2.4.2 Solid Wall Panels

(a)

voids and solids. This trend has resulted in some ingenious precast concrete configurations with the use of
L- and T-shaped panels to reduce the number of cost
increasing joints. These panel shapes also allow efficient erection and thus reduce installation cost. Some
typical panel arrangements are shown in Fig. 2.4.5.

2.4.2 Solid Wall Panels
Solid wall panels use finish, shape, size, and repetition as the major design and cost considerations. The
high level of design flexibility possible with custom wall

Fig. 2.4.5 Typical arrangement of precast concrete panels.

(b)

(b)
(a)

(c)

(d)

(f)

(g)

(e)

Fig. 2.4.4(a) & (b)
Sybase Information Connect Division
Boulder, Colorado;
Architect: OZ Architecture; Photos: OZ Architecture.

64

|

ARCHITECTURAL PRECAST CONCRETE

Individual panel

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.2 Solid Wall Panels / 2.4.3 Window Wall Panels

panels allows for a wide variety of architectural appearances. The precast concrete cladding in Fig. 2.4.6 was
articulated by the use of horizontal and vertical rustications to reduce the scale of the massive, windowless
walls of a courthouse.

They are either one story in height and made as wide as
possible, or cast narrower to span vertically for two to
three floors, except in high seismic areas where story drift
may control design. Window openings placed within the
body of the panel provide closer tolerances for window

2.4.3 W
 indow Wall Panels

Fig. 2.4.6
Charles Evans Whittaker Courthouse, Kansas City, Missouri;
Architect: Ellerbe Becket/Abend Singleton Associated Architects;
Photo: Timothy Hursley.

Window wall panels may be flat or heavily sculptured.
They may contain a single opening or series of windows.

ARCHITECTURAL PRECAST CONCRETE

|

2

65

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.3 Window Wall Panels / 2.4.4 Spandrel Panels

installation than when the window is defined by the edges of separate spandrels
and mullions. The project in Fig. 2.4.7(a)
and (b) uses mostly window box units that
typically span column to column. The large
panel size is disguised by the use of both
horizontal and vertical reveals.

(b)

2.4.4 S pandrel Panels
Spandrel panels are horizontal units
that separate adjacent strips of glass.
They may be cast flat, have returns at the
top and/or bottom, or be heavily sculpted. A designer will sometimes require
that the structural frame of a building
be expressed in the building’s façade. In
(a)

Fig. 2.4.7(a) & (b)
116 Huntington Avenue
Boston, Massachusetts;
Architect: CBT/Childs Bertman Tseckares Inc.;
Photos: Wayne Soverns Jr.

66

|

ARCHITECTURAL PRECAST CONCRETE

such cases, the use of precast concrete spandrel elements, made up either as a series of individual units or
as one unit extending between columns with support
located on the floor or on the column, is an aesthetically appropriate solution. Emphasizing the linearity or
horizontality of the design (Fig. 2.4.8), white spandrel
panels wrap the building at each floor interrupted at
the middle by the four-story entry. The deep-ribbed
spandrels in Fig. 2.4.9(a) have a unique profile that

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.4 Spandrel Panels

2

Fig. 2.4.8
Liberty Property at Huntington Square, Miramar, Florida;
Architect: Retzsch Lanao Caycedo Architects;
Photo: RLC Architects, P.A.

Fig. 2.4.9(a) & (b)
Oak Brook Pointe Office Center
Oak Brook, Illinois;
Architect: Wright Architects Ltd.; Photo: Hedrich Blessing.

(b)

cants outward five degrees from vertical (Fig. 2.4.9[b]).
Each panel has two finishes of a light gray concrete,
acid-etched at the flutes and lightly sandblasted at the
top, creating the look of honed and flamed salt-andpepper granite. At the stepped end walls, the windows
follow the same tilt as the spandrels, creating a continuous projection. Tilting the windows required creation
of a spandrel panel that was hung on an angle. In contrast, the typical spandrel rests normally on the slab
but has a canted face. These two sections join gracefully and seamlessly at each of the building’s twelve
corners. Special corner pieces, formed with sequential
casts rather than in a V-mold, provided total continuity between elevations. The result is that the deep rib
reveals are in perfect alignment along the entire perimeter of each floor.

ARCHITECTURAL PRECAST CONCRETE

|

67

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.5 Column Covers and Mullions

(b)

(a)

Fig. 2.4.10(a) & (b)
901 New York Avenue
Office Building; Washington, DC;
Architect: Davis, Carter, Scott;
Photos: Eric Taylor Photography,
© Eric Taylor Photo.com.

2.4.5 C olumn Covers and Mullions
Column covers and mullions are usually a major focal
point in a structure. These units may be broad or barely
wider than the column itself and run vertically up a
structure. They are often used to conceal the structural columns and may completely surround them at
the ground level. Column covers are usually manufactured in single-story units and extend either from floor
to floor or between spandrels; however, units two or
more stories in height may be used. In order to minimize erection costs and horizontal joints, it is desirable
to make mullions as long as possible, subject to limitations imposed by weight and handling. Also, in many
cases it may be desirable to combine the column cover
or mullion with adjacent spandrel to minimize joints.
The column covers in Fig. 2.4.10(a) and (b) are the
major focal areas of the building. The buff precast
concrete units were given an acid-etched finish. In Fig.
2.4.11, the precast concrete units are used as beam
and column covers in the interior to visually integrate
the exterior and the atrium. In Fig. 2.4.12, the precast

68

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 2.4.11
Newton Pavilion at Boston Medical Center, Boston, Massachusetts;
Architect: SMMA Hoskins Scott formerly Hoskins Scott Taylor &
Partners Inc.; Photo: Paul Gobeil Photography.

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.4.5 Column Covers and Mullions / 2.4.6 Wall-Supporting Units

2

Fig. 2.4.12
Palisades Office Complex
Atlanta, Georgia;
Architect: Cooper Carry and
Associates;
Photo: Cooper Carry Inc.

concrete column covers provide an appealing accent to
the entrance. In Fig. 2.4.13, the precast concrete mullions spaced 5 ft (1.5 m) on center lend verticality to
the 20-story tower surfaces and emphasize the desired
delicacy of scale and finish.

2.4.6 W
 all-Supporting Units
Wall-supporting units are precast concrete cladding
units that support a portion of the wall, but carry no
loads from floors or roof slabs. These units cannot
be removed from a wall without affecting the stability of other units and are normally designed so that
their largest dimension is vertical, although they may
be horizontal.
Where possible, the lower two to three stories of
panels should be “stacked” to support the wall above
Fig. 2.4.13
Eagle Gate Plaza & Office Towers
Salt Lake City, Utah;
Architect: Cooper Carlson Duy Ritchie, Inc.;
Photo: Rodriguez & Associates, LC.

ARCHITECTURAL PRECAST CONCRETE

|

69

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.6 Wall-Supporting Units

them up to the roof level, or any portion of this height,
and supported directly on the foundation wall to eliminate load support connections and speed erection (Fig.
2.4.14). These units may be quite slender and support
considerable wall height if they can be tied into the
structure as required for lateral stability. Panels have
been successfully stacked higher than three stories, but
these cases require engineering considerations regarding the impact of thermal movements, lateral building
drift, and structure deflections.

concrete panel with 2 in. (50 mm) of German limestone. There is 60,000 ft2 (5600 m2) of cladding with
a 2 ft (0.6 m) return. This profile would have made
the connection system very complex if the panels were
individually gravity-load-supported in lieu of stackloaded. The simplicity of the stacked system allowed the
use of a slotted insert and strap type tieback connection throughout.
Precast concrete units designed to carry their own
weight over a considerable width are also considered
wall-supporting units. This width may be equal to the
column spacing for the exterior wall or a multiple thereof. Where such units are bearing at, or in close proximity to, the columns, edge beams may be eliminated. If
edge beams are needed for other structural reasons
they may afford savings in size and/or reinforcement in
the panels. Because of size or weight limitations, such
units are normally made only one story high so that
the width is the largest dimension. When the panel
spans across several columns, the potential for deflection (or rotation) of the edge beams caused by the

Wall-supporting units can be made in one piece
through several stories. The building frame carries
only lateral loads from the precast concrete panels, as
all axial loads from the wall panels, are supported by
the foundation. This reduces the need for larger steel
members around the perimeter of the building, resulting in a more economical steel superstructure. The
weight of a group of stacked units should all be carried by a single designated floor. Erection techniques
to accommodate predetermined partial load distribution between floors are not economically feasible. It
has proven to be impractical
to support adjacent units on
Fig. 2.4.14(a) & (b) Stacked and individually supported panels.
alternate floors. The recommended practice is that a
specific floor be designated
Real joint
and designed to take the
load of all precast concrete
units passing it.
Wall-supporting units may
answer a particular design
consideration for structures
where the exterior columns
are set back from the edge
of the floor slab. A cantilevered floor will deflect to a
certain degree over a period
of time due to the weight
of the wall units. By using
stacked units for this condition, the designer can respond to this consideration.
The art institute in Fig.
2.4.15 is a four-story, stackloaded cladding project with
a very pronounced profile.
The cladding consists of a 6
in. (150 mm) thick precast

70

|

ARCHITECTURAL PRECAST CONCRETE

Real joint

False joint
(if required)

False joint
(if required)

Real joint

Real joint

Solid shim

False joint

Stacked panels
(a)

False joint

Individually supported panels
(b)

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.6 Wall-Supporting Units

2

wall, any long term deflection in
the unit will normally have taken
place by the time of erection.
The Phillies wanted the exterior
of its ballpark to be consistent with
the city’s dominant architectural
elements, so brick was the natural
choice for the cladding. To meet
the budget and schedule, the design and construction teams used
architectural precast concrete with
inset brick (Fig. 2.4.16). The precast
concrete panels were designed to
be self-supporting elements, along
the ramps and at the arcades at the
south and west faces of the office
buildings (along the pedestrian entrances). The piers that frame the
arcade were shipped as U-shaped
column sections and support the
weight of the infill spandrel panels located above the windows
and arcade openings. This reduced
the amount of weight to be supported by the steel structure and
Fig. 2.4.15
Target Wing – Minneapolis Institute
of Arts
Minneapolis, Minnesota;
Architect: RSP Architects.

weight of wall units is reduced. Where
several units are carried by one beam,
resulting deflection may create tapered
joints and the possible touching of units
at their tops. Designing the panel to span
the distance between the columns with
one unit provides a deep beam and, consequently, much less deflection. By storing and supporting such units in a way
similar to their ultimate position in the

Fig. 2.4.16
Phillies Ballpark
Philadelphia, Pennsylvania;
Architect: HOK Sports Facilities Group, LLC; Ewing
Cole/Cherry Brott, joint venture.

ARCHITECTURAL PRECAST CONCRETE

|

71

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.4.6 Wall-Supporting Units

Fig. 2.4.17
Franklin High-Tech Center, Franklin Township, New Jersey; Architect: Herbert Beckhard Frank Richlan
& Associates; and Brandt-Kuybida, joint venture; Photo: O. Baitz Inc. Photography of Architecture.

eliminated the need for supplemental steel members
to support and brace hand-laid masonry. These panels
were constructed with articulations, brick patterns, and
granite accent bands that hark back to older ballparks
and complement the city’s architecture.
The five buildings of the
industrial/office complex
in Fig. 2.4.17 have strongly horizontal, acid-etched,
light gray, ribbed precast
concrete panels. The majority of the panels are 8 ft
(2.4 m) high, 20 ft (6.1 m)
wide, and 8 in. (200 mm)
thick. The ribs provide the
necessary structural stiffening, making the panels
self-supporting and stackable. By stacking each
panel on the panel below,
the gravity loads from the
panels are carried directly
to the foundation and do
not introduce additional
load to the superstructure.
The ribs also act as stiffeners for the stresses during
handling and erection.

72

|

Having a self-supporting (stacked) precast concrete system saved engineering time and helped meet a very
tight schedule (Fig. 2.4.18). The project had over 1100
panels and 113,000 ft2 (10,500 m2) of 6 in. (150 mm)
thick cladding. A gray color was used with four differ-

Fig. 2.4.18
Minnesota Department of Revenue Building
St. Paul, Minnesota;
Architect: HGA Architects
and Engineers.

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.4.6 Wall-Supporting Units / 2.5 Loadbearing Wall Panels or Spandrels-General

2

Fig. 2.4.19(a), (b) & (c)
Four Gateway Plaza
Colorado Springs,
Colorado;
Architect: DCA Architects.

(a)

(b)

(c)

ent finishes. Additionally, buff-colored precast concrete
inlays accent the gray.
The three-story structural steel office building in Fig.
2.4.19(a) is supported on the exterior by low profile
precast concrete spandrels and column units that allow
for large windows offering views of Pike’s Peak. The
walls are stackable and the precast concrete was erected in two phases (Fig. 2.4.19[b]). The two-story exterior walls and spandrels were erected first and braced
to the first-floor steel structure. Then the second-floor
steel was placed on the walls and spandrels. After the
topping was placed and cured on the second floor, the
third-floor columns and spandrels were erected and
the braced to the second floor. The architectural walls
and spandrels consist of a two-color, acid-etched finish with a form liner used to create a chiseled stone
band, while reveals give a human scale to the project
(Fig. 2.4.19[c]).

2.5 LOADBEARING WALL PANELS
OR SPANDRELS
2.5.1 General
Loadbearing façades have both an aesthetic and a
structural function. In building practice, the most economical application of architectural precast concrete is
as loadbearing structural elements. Loadbearing units
become an integral part of the structure, taking the
vertical and horizontal floor and roof loads and/or
transferring horizontal loads into shear walls or service
cores. Such an arrangement can be economical, not
only from a structural design standpoint, but also from
the viewpoint of overall construction. In some cases,
the loadbearing elements also can contribute to the
horizontal stability of the building. Each loadbearing
element plays an essential role in the structural integrity or stability of the building.

ARCHITECTURAL PRECAST CONCRETE

|

73

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5 Loadbearing Wall Panels or Spandrels-General

Architectural precast concrete cladding is noted for
its diversity of expression, as well as its desirable thermal, acoustic, and fire-resistant properties. Commonly
overlooked is the fact that concrete elements normally
used for cladding applications, such as solid wall panels, window wall, or spandrel panels, have considerable inherent structural capabilities.

columns can be substantial and, depending on the floor
plan, partition layout flexibility can be enhanced. Also,
unlike a steel frame, a loadbearing precast concrete system eliminates the need for cementitious fireproofing
and the associated costs and schedule impact of tenting
and temporary heat required for applicaton and curing
of fireproofing material.

In the case of low- or mid-rise structures, the amount
of reinforcement required to handle and erect a precast
concrete component is often more than that necessary
for carrying design imposed loads. Thus, with relatively
few modifications, many cladding panels can function
as loadbearing members. For taller buildings, additional
reinforcement may be necessary for lower-level panels.

Structural Depth. One of the perceived disadvantages
of a loadbearing precast concrete system, compared to
a steel structure, is the potential larger structural depth
required for a given span and the resulting increase in
floor-to-floor height. Taller floors increase the amount of
exterior wall area required to enclose a given floor plate,
which translates directly into increased construction
costs for a structure’s skin. However, an understanding
by the designer of the scope of such increases and strategies to minimize their impact through efficient floor
plate design can mitigate this disadvantage.

The slight increases in loadbearing wall panel cost (due
to reinforcement and connection requirements) can
usually be more than offset by the elimination of a separate perimeter structural frame. Depending upon the
application, the loadbearing panels also may reduce or
eliminate a structural core or interior shear walls, particularly in buildings with a large ratio of wall-to-floor area.
The increase in interior floor space gained by eliminating

A typical floor-to-floor dimension for a 45 ft (13.7 m)
clear-span steel structure with an 18-in.-wide (460 mm)
flange beam averages about 13 ft 4 in. (4 m). This distance accounts for a 9 ft (2.7 m) finished ceiling height

Fig. 2.5.1 Exterior wall-to-floor ratio.
Building

74

|

Number
of Floors

Floor
Plate Area

Perimeter
Wall
Floor Area

Gross
Area

Net Area

Net Area
Gross Area

Walking Distance

Corridor
Area

Corner to Core

Diagonal

A

4

40,000

0.32

160,000

150,200

93.40%

5900

225

430

B

4

41,300

0.34

165,400

165,400

94.00%

7200

255

510

C

4

40,100

0.32

160,600

160,600

95.00%

5700

235

455

D

4

39,300

0.33

159,200

159,200

95.00%

7200

175

330

E

4

40,000

0.32

160,000

154,300

96.00%

5700

180

330

F

2

80,500

0.27

161,000

155,300

96.50%

7900

330

620

G

3

52,100

0.29

156,400

149,000

95.50%

9000

250

400

H

4

40,000

0.31

160,100

152,500

95.30%

9200

270

315

I

4

39,900

0.33

156,800

147,700

94.20%

5100

170

240

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5 Loadbearing Wall Panels or Spandrels-General

2

Fig. 2.5.2 Schedule for loadbearing precast concrete project.
SCHEDULE
Week #

1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19

20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56

EVENT/TASK
SCHEMATIC DESIGN - SD
OWNER REVIEW
DESIGN DEVELOPMENT - DD
CONSTUCTION DOCUMENT - CD
CONSTRUCTION ADMIN - CA
RFP / CM / GC SELECTION
PRE-CONSTRUCTION
CONSTRUCTION PHASE
TOTAL PRECAST
Work Sessions - Precast SD
Commitment to Precaster
Precast Design Assist
Precast Design and Detailing
Plant Fabrication
Erection

Note: Schedule durations are for illustration only.

and the required clearances for fireproofing, lighting, and
ductwork. For a similar span in a loadbearing precast concrete system using 24-in.-deep (0.6 m) double tees, the
required floor-to-floor dimension is 14 ft 0 in. (4.3 m).

The difference between these building shapes with varying
numbers of floors for a program of 160,000 ft2 (14,860 m2)
is shown in Fig. 2.5.1, and it amounts to a 26% swing in
the wall area required to enclose identical areas.

The extra 8 in. (200 mm) of exterior wall required for a
mid- to high-rise structure represents a 5% increase in
wall area over the steel system. But that increase must
be put into perspective with other aspects of the design,
especially floor-plate geometry and floor vibration.

The lesson to be learned from this is that floor-plate
geometry and building height have much more of an
impact on exterior wall area than a potential 5% penalty due to the increased floor-to-floor height required
for a total precast concrete system.

In some areas of the country 20 in. (500 mm) deep
double tees are used for a 45 ft (13.7 m) span, while
steel designers may use, a 24 in. (600 mm) wide flange
beam ito minimize floor vibrations. Precast concrete
has a much stiffer floor system than steel. When this
construction configuration is used the floor-to-floor
dimensional difference between precast concrete and
steel is eliminated.

To realize the full potential of loadbearing unit usage with no sacrifice in aesthetic advantages, both the
engineer of record and precast design engineer should
be involved from the initial concept stage of the project. Considerations should include the load effects on
member dimensions, coordination of temporary bracing, connections, and erection sequencing.

Exterior Wall-to-Floor Ratio. An important consideration when determining the efficiency of a particular
structure is the ratio of exterior wall to floor area. The
lower this ratio, the more efficiently a particular floor
plan shape encloses interior space and reduces the required amount of exterior wall.
As we know, a circular shape provides the most efficient
perimeter defining a given area. Floor plan shapes that
most closely resemble a circle (such as a square as opposed
to long and thin rectangles, T-or L shapes) offer better
floor-to-wall area ratios. Also, for buildings of a given total
square footage those that have fewer floors have better
wall to floor area ratios than those that have more floors.

To take maximum advantage of loadbearing and wallsupporting units, the decisions as to their functions
should be made before structural design of the building
frame has progressed to a stage where revisions become
costly for a given project schedule. An earlier award of
the precast concrete contract will be necessary when using loadbearing units than for non-loadbearing façades
(Fig. 2.5.2). Cost savings tend to be greatest in low- to
mid-rise structures of three to ten stories. As with all
precast concrete applications, further economies can be
realized if the panels are repetitive. Besides minimizing
the number of molds necessary, repetitive panel designs
enable repetitive connections.
The use of a loadbearing precast concrete system can

ARCHITECTURAL PRECAST CONCRETE

|

75

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5 Loadbearing Wall Panels or Spandrels-General

(a)

(b)

Fig. 2.5.3(a) & (b)
NBSC Headquarters Building
Greenville, South Carolina;
Architect: Neal + Prince & Partners;
Photo: (a) Fred Martin Jr./Fred
Martin Photography Inc.

trim two-and-a-half to three months off the typical construction schedule for a mid-sized, 100,000 ft2 (9300
m2) structure. An example of the speed of construction is shown for the office building in Fig. 2.5.3(a).
The erection phase of this four-story, 75,000 ft2 (7000
m2) Class A office building was completed in 44 days
(Fig. 2.5.3[b] and [c]. The project incorporates many
precast concrete components from exterior bearing
walls with integrally cast thin brick, columns, inverted
tee beams, double-tee floor and roof members, and
interior shear walls.

76

|

ARCHITECTURAL PRECAST CONCRETE

Architectural loadbearing panels can be used effectively to renovate and rehabilitate old deteriorated structures. These panels can be used not only in all precast
concrete structures but also in structural steel-framed
structures and cast-in-place concrete structures.
Design guidance for using loadbearing architectural
precast concrete wall panels can be found in Sections
4.2.5 and 4.2.7, as well as the PCI Design Handbook
– Precast and Prestressed Concrete, MNL-120.

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5 Loadbearing Wall Panels or Spandrels-General

2

Fig. 2.5.3(c) Construction sequence.

Precast Day 1 — Foundations ready, first piece arrives.

Precast Day 7 — Building begins taking shape.

Precast Day 15 — Interior shear walls.

Precast Day 25 — Shear structure.

Precast Day 37 — Nearing completion.

Precast Day 44 — Structurally and architecturally completed.

ARCHITECTURAL PRECAST CONCRETE

|

77

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.2 Shapes and Sizes

2.5.2 S hapes and Sizes

(a)

Architectural loadbearing components can be provided in a variety of custom-designed or standard section shapes. A wall system can be comprised of flat
or curved panels (solid or insulated) (Figs. 2.5.4[a] and
[e]), window or mullion panels (Figs. 2.5.4[b] and [c]),
or ribbed panels (Fig. 2.5.4[d]). Each type of panel will
readily accommodate openings for doors and windows.
Figures 2.5.4(b), (c), and (d) illustrate various types of
ribbed panels. The panel shown in Fig. 2.5.4(c) is a horizontal Vierendeel truss window mullion panel, while
the other panels are vertical window mullion panels.
Figure 2.5.4(e) shows an exterior horizontal spandrel
that would be used as part of a column-wall system.
In the interest of both economy and function, precast
concrete panels should be as large as practical, while
considering production efficiency and transportation
and erection limitations. By making panels as large as
possible, numerous economies are realized: the number of panels needed is reduced, fewer joints (waterproofing requirements) and connections are required,
and the erection cost is lower.
Panels may be designed for use in either vertical or
horizontal positions. For low-rise buildings, complex
connection details can be minimized by spanning loadbearing panels vertically through several stories; consequently, the economic advantages of loadbearing wall
panels are increased. For high-rise buildings, it is normally more practical to work with single-story horizontal panels connected at each floor level. The elements
can be more slender, simplifying the erection.
The 16-story, 249,000 ft2 (23,200 m2) office building is
topped out at 248 ft (75.6 m) above grade (Fig. 2.5.5[a]).
It has single-story, horizontal loadbearing panels that are

(a) Flat or
(A)
insulated
panel

(b) Vertical window or
mullion (B)
panel

(c) Horizontal window
(C)
or mullion
panel

Fig. 2.5.4 Various types of architectural loadbearing wall panels.

78

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 2.5.5(a) & (b)
United Bank Tower, Colorado Springs, Colorado;
Architect: Klipp Partnership, P.C.

typically 14 ft 6 in. x 16 ft (4.4 x 4.9 m) by 8 in. (200
mm) thick. Vertical load transfer at the exterior of the
building was accomplished by spanning from the building core to the exterior walls horizontally to monolithically cast 13 x 30 in. (330 x 760 mm) column elements

(d) Ribbed panel

(D)

(e) Horizontal spandrel as
(E) system
part of a column-wall

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.2 Shapes and Sizes

2

although, multistory panels 45 ft (13.7 m)
in height have been fabricated and delivered. Panels should be designed in specific
widths to suit the building’s modular planning. When such a building is designed to
take the best advantage of modularity, the
economic advantages of loadbearing wall
panels are significantly increased.

(b) Column free interior.

within the wall panels. The monolithic columns reduced the number of components to be erected and
the subsequent connections. Also, 11 x 15.5 in. (280 x
390 mm) precast concrete mullions, integral with the
panels, allowed incorporation of four windows in each
panel. A column free interior allows for flexible use of
office space (Fig. 2.5.5[b]).
The maximum panel size that can be transported
is impacted by local conditions, such as bridge and
overhead utility clearances, site access, and regulatory agencies such as State and Federal Departments
of Transportation. In general, a panel up to 12 ft (3.7
m) tall and 30 ft (9.1 m) long is a manageable size;

Load uniformity is one of the important
advantages for high-rise, loadbearing panel structures. This approach can produce
evenly distributed loads on the perimeter
foundations and reduces the tendency for
differential settlement. The jointed nature of
the façade also makes it more tolerant of
any differential settlement that may occur.
Curves are easily created with precast concrete. On
curved panels, a continuous supporting ledge cast on
the inside face is preferred to provide bearing for floor
and roof members and to stiffen the panels to minimize warping. Figure 2.5.6 shows the skeleton of various components at the right, and an ever-more finished
look of the structure to the left. At far left is how the
finished structure will appear.
Flat and curved molds, each 13 ft (4.0 m) high and
30 ft (9.1 m) wide, produced the loadbearing panels for the eight-story office building in Fig. 2.5.7(a).

Fig. 2.5.6 Schematic of loadbearing precast concrete building.

ARCHITECTURAL PRECAST CONCRETE

|

79

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.2 Shapes and Sizes

Individual, non-repetitive stone
textures comprise the horizontal
moldings that encircle the building at every level. To produce
these textures, the architects
chose stone from a local quarry
and worked with stone cutters
to develop 70 different profiles.
A urethane mold was then fabricated from each stone. The
flexible urethane casts of the
individually chiseled limestone
blocks were then placed within
the larger panel molds in a variety of patterns, adding depth,
detail, and a non-repetitive
quality. The process of casting a
highly articulated, buff-colored,
stone-textured loadbearing wall
solved structural and cost constraints and avoided the process

(b)

(a)

of attaching the stone to the precast concrete panels
(Fig. 2.5.7[b]).

Fig. 2.5.7(a) & (b)
198 Inverness Drive West
Englewood, Colorado;
Architect: Pouw & Associates, Inc.;
Photos: Pouw & Associates, Inc.

80

|

ARCHITECTURAL PRECAST CONCRETE

Wall panel size and shape can be affected by the details and locations of the vertical and horizontal panelto-panel connections. Gravity load transfer between
panels, gravity and axial load combinations caused by
lateral loadings, or size of window openings can become major factors influencing panel structural dimensions and connection design. For most precast concrete

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.2 Shapes and Sizes / 2.5.3 Design Considerations

exterior bearing wall structures, the gravity dead and
live load condition will control structural dimensions.

through tension connections and high-strength grout
in horizontal joints.

When stemmed floor members, such as double tees,
are used, the width of loadbearing walls or spandrels
should module with the double-tee width. For example, for 12 ft (3.6 m) double tees, walls should be 12,
24, or 36 ft (3.7, 7.3, or 11 m) wide. Local precast
concrete producers should be contacted to determine
their particular module.

As in all precast concrete construction, the transfer of
vertical load from element to element is a major consideration. Differences in section shape, architectural
feature, and unit stress result in a variety of solutions
and types of connections.

Inverted tee beams typically are used on interior
spans. To minimize floor-to-floor dimensions, double
tees are frequently dapped at interior beam lines and
at exterior spandrels. Dapping is generally not necessary on vertical wall panels.

2.5.3 D
 esign Considerations

(See Sections 4.2.5 and 4.2.7 also)

In recent years, tremendous advances have been made
in precast concrete structural engineering technology.
Greater knowledge regarding connections and wall
panel design has made it possible to use architectural
loadbearing precast concrete wall panels more cost effectively. Solid panels, or panels with small openings,
constitute true bearing walls because they are primarily
stressed in compression. With solid flat panels, load
path locations can be determined easily.
As openings in the wall become larger, loadbearing
concrete panels may approach frames in appearance
and the concentration of load in the narrower vertical sections increases. In multistory structures this load
accumulates, generally requiring reinforcement of the
wall section as a column (at panel ends and at mullions
between windows) designed for biaxial bending due
to load eccentricities.
Loadbearing panels and shear walls, generally, will
be supported continuously along their lower surface.
Continuous footings, isolated piers, and grade beams
or transfer girders may support them. Lower floors can
be framed with beams and columns, to allow for more
open space on these levels, while the structural system
on the upper floors can consist of bearing walls.
When this is done, careful attention must be paid to
the effective transfer of the lateral forces to the foundation. As with a vertical irregularity in any building in
a seismic zone, the structural engineer should make
a careful assessment of the behavior and detailing. In
multistory bearing walls, design forces are transmitted

2

Depending on wall section and foundation conditions, a loadbearing wall panel can be fixed at the base
(shear walls for lateral forces) with the roof elements
freely supported on the panel. Alternately, depending
on the shape of the building, wall element flexural
stresses can be reduced by providing pin connections
at the foundation and shear wall bracings at the ends
or across the building to ensure lateral stability.
Loadbearing or shear wall units should be the primary
design consideration if one or more of the following
three conditions exist:
1. There is inherent structural capability of the units
due to either their configuration or to sufficient panel thickness. The sculptural configuration of units
often enables them to carry vertical loads with only
a slight increase in reinforcement. For example, the
precast concrete units may have ribs or projections
that enable them to function as column elements
for the structure. Ribs may be part of the architectural expression or, where flat exposed surfaces are
required, may be added to the back of panels for
additional stiffness. Projections do not have to be
continuous or straight, as long as no weak point
is created within the units. Generally, there is little
cost premium for sculptured panels when there is
adequate repetition. Similarly, some flat panels (including sandwich wall panels) may be sufficiently
thick to carry loads with only minor increases in reinforcement. Structural design of panels with insulation between layers of concrete (sandwich panels)
usually ignores the loadbearing capacity of the nonbearing wythe,unless designed as composite panels.
If possible, the structural wythe of a sandwich wall
panel should be kept on the temperature-stabilized
side of the building to reduce thermal stresses due
to temperature variation.
2. A uniform structural layout of the building facilitates
favorable distribution of lateral forces from wind or
earthquake loads. Plus, this uniformity lends itself to
repetitive, economic castings (Fig. 2.5.8). This concept
is difficult to employ if the load paths are continually

ARCHITECTURAL PRECAST CONCRETE

|

81

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.3 Design Considerations

changing from floor to floor. Cast-in-place topping on
precast concrete floor units enables the floors to act
as diaphragms, distributing lateral forces and reducing
both individual wall unit and connection loads.
3. The building has a central core or bay designed to
resist lateral forces and transfer them to the foundation (Fig. 2.5.9). When the core creates a tor-

sional irregularity, it should be supplemented with
other lateral–force resisting elements either in the
interior or on the exterior. Plan irregularities created by the extended wings of a C or Z shape, are
particularly problematic in moderate or high seismic risk areas. Because the core or bay provides
structural rigidity, panel-to-floor connections can

Maroon

Crescent IV

Aurora Municipal Center
TCI

Starz Encore

Fig. 2.5.8 Floor plates of loadbearing buildings by Barber Architecture.

82

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.3 Design Considerations

remain relatively simple. A typical building core
may contain an elevator lobby, elevators, stairways, mechanical and electrical equipment, and
space for air ducts. While the core is being erected
or cast, the precaster can proceed with the fabrication of the exterior wall units and install them as
the shear wall or core is being constructed, often
saving construction time.
Loadbearing wall panels used to construct the
building core are connected after erection to form
composite T-, L-, U-, or boxed-shaped sections in
plans. The main advantages of precast concrete
cores versus cast-in-place cores are surface finish
quality, faster construction, and greater flexibility
of the precast concrete erection sequencing.
Loadbearing spandrel panels are essentially perimeter
beams that may extend both above and below the floor
surface, and which transfer vertical loads from the floor
or roof to the columns. Loadbearing spandrels are either
ledged, pocketed, or have individual or button haunches
(also known as spot corbels) to support floor and/or roof
members. Steel shapes and plates may be cast into the
panels to reduce haunch height and, therefore, floor-tofloor height. Non-loadbearing (closure) spandrel panels
may have much the same cross-section as loadbearing
spandrels without ledges, pockets, or haunches.
Precast concrete building elements are commonly reinforced with welded wire reinforcement, mild steel
reinforcement, or prestressing steel. Unless analysis or
experience indicates otherwise, both loadbearing and
non-loadbearing panels should be reinforced with the
amount of steel reinforcement specified in the appropriate building code.
Lateral loads applied perpendicular to the wall are the
result of wind or seismic forces, and are usually transmitted to vertical stiffening cores, shear walls, structural frames, or other stabilizing components by roof
and floor members acting as horizontal diaphragms.
This reduces the load on individual wall units and their
connections (Fig. 2.5.10). The connections between
façade elements and floor members are normally designed as hinges in the direction perpendicular to their
plane.
Vertical continuity is achieved by providing connections at horizontal joints of vertical members. Columns
should be braced at each level through a continuous
load path to the diaphragm.

(a)

(e)

(i)

(b)

(c)

(f)

2

(d)

(g)

(h)

(j)

(k)

Fig. 2.5.9 Plan view of possible locations of vertical cores with
respect to loadbearing walls. (Note: Although the building core
is an important element of the lateral force resisting system, it
may be insufficient to handle the torsional effects of eccentrically applied loads in some of these plans. Also some plans
have re-entrant corners that create plan irregularties.)

The all–precast concrete structure with architectural precast concrete loadbearing wall panels and
precast concrete core walls, beams, and double
tees provided the neoclassic look of traditional ma-

Load

Façade

Floor
Units
Plan

Fig. 2.5.10 Principle of diaphragm action in precast
concrete floors and roofs.

ARCHITECTURAL PRECAST CONCRETE

|

83

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.3 Design Considerations

(a)

(b)

Fig. 2.5.11(a), (b) & (c)
Aurora Municipal Center
Aurora, Colorado;
Architect: Barber Architecture;
Photo: (a) Michael Peck.
(c)

84

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.3 Design Considerations

sonry used by many historic city halls (Fig. 2.5.11[a]).
Two five-story office wings feature acid-etched precast concrete panels with punched windows and
deep reveals. The building’s wings are connected
by a central, six-story curving section to provide
a 56,000 ft2 (5200 m2) floor plate(Fig. 2.5.11[b]). The
massive architectural columns and spandrels at the second level support all the loadbearing, two-story architectural walls above. Three cranes working simultaneously
erected the building as separate sections or “towers”
because of an accelerated schedule—this allowed 55
pieces to be erected in a single day (Fig. 2.5.11[c]). The
result was significant schedule advantages over a steel frame structure with
brick or stone cladding.

2

components and one spanning arch element comprise
each of these typical bays (Fig. 2.5.12[b]). The entry
drum is comprised of five bays, each 19 ft (5.8 m)
wide, constructed in a similar manner.
A modern interpretation of a neo-classical
Mediterranean style, the mixed-use structure in Fig.
2.5.13(a) combines retail on the first floor with four
floors of parking above. The remaining four floors
consist of office spaces and luxury condominiums.
Floridian colors helped to anchor the project to its locale. The structural concept was originally designed
(b)

The 20,000 ft2 (1900 m2) bank building integrates structural clarity with the
classical form of traditional bank architecture (Fig. 2.5.12[a]). The building is
rendered in modular loadbearing precast concrete panels. The main body of
the building is constructed of 18 typical
bays, each 15 ft (4.6 m) wide and 30 ft
(9.1 m) high. Two loadbearing column

(a)

Fig. 2.5.12(a) & (b)
Park National Bank, Mt. Prospect, Illinois;
Architect: Cordogan, Clark & Associates;
Photos: Cordogan, Clark & Associates.

ARCHITECTURAL PRECAST CONCRETE

|

85

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.3 Design Considerations

Fig. 2.5.13(a) & (b)
The Metropolitan
Jacksonville, Florida;
Architect: KBJ Architects Inc.;
Photos: Brian Griffis.

(a)

(b)

to use cast-in-place concrete and was converted to loadbearing precast concrete to accelerate the schedule (Fig. 2.5.13[b]). Precast
concrete also made sense because of an extremely restrictive building site. The building
footprint is 120 x 240 ft (37 x 73 m) and the
lot size is 125 x 300 ft (38 x 91 m).
Two-story, loadbearing, punched window
wall units, 14 x 31 ft (4.3 x 0.95 m) in section
and 9 1/2 in. (240 mm) thick, were stacked
and engineered to become an integral part
of the structure taking the vertical and horizontal floor and roof loads (Fig. 2.5.14 [a]
and [b]). The developer was able to eliminate

86

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.3 Design Considerations

the perimeter structural steel columns. Also, the precast concrete elevator shaft was designed as a lateral
load resisting frame. From an architectural viewpoint,
the flexibility of the precast concrete system offered
the owner unlimited colors and textures, deep reveals,
custom designed cornice details, and a virtually main-

2

tenance-free exterior finish. The precast concrete system for the 4-story, 117,000 ft2 (10,900 m2) Class A
office building saved more than $250,000 over a conventional precast concrete and structural steel building
of similar size.

(b)

Photo:
Brian Griffis.

(a)

Fig. 2.5.14(a) & (b)
Deerwood North 300, Jacksonville, Florida; Architect: Rolland Delvalle & Bradley; Photo: Dennis O’Kane.

ARCHITECTURAL PRECAST CONCRETE

|

87

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.3 Design Considerations

Fig. 2.5.15(a), (b) & (c)
Arapahoe County Centrepoint Plaza, Aurora, Colorado; Architect: Barker Rinker Seacat Architecture, P.C.;
Photos: Barker Rinker Seacat Architecture.

(a)

(c)

band also has a feather feature created with a form
liner. The exterior columns are colored concrete with
an acid-etch finish and the top of the columns has a
medallion of feathers cast with a form liner. The entrance of the building features large wall panels with a
form liner finish of large feathers. The feather theme is
to honor and respect the Arapahoe Native Americans
(Fig. 2.5.15[c]).

The four-story total precast concrete structure in
Fig. 2.5.15(a) has 175,000 ft2 (16,300 m2) of double
tees supported by precast concrete
beams, architectural precast concrete walls, and one-piece, fourstory columns. The erection of the
precast concrete was completed in
nine weeks (Fig. 2.5.15[b]). A design feature of the structure is that
the second floor extends out from
the basic line of the building. This
required a large transfer beam to
support the upper exterior wall back
at the building line. The spandrels
use thin inset brick framed top and
bottom with colored concrete that
has an acid-etched finish. The top

88

|

ARCHITECTURAL PRECAST CONCRETE

Aesthetics were the driving force behind the innovative precast concrete techniques used for the headquarters building in Fig. 2.5.16(a). For while the use
of an all-precast concrete structural system with loadbearing exterior panels saved time and money and provided a variety of design advantages, the most striking
aspect of the project was the dramatic image created
by the precast concrete panels, which include a random cut-stone design on the lower levels. Some of the
(b)

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.3 Design Considerations

base panels are 22 in. (560 mm) thick and weigh up to
58,000 lb (26,300 kg) each. The building’s heavily rusticated base needed to match the rough texture and
color tone of the thermal granite finish on the adjacent
headquarters building. Because repetition becomes
obvious when panels are placed next to each other,
nearly 280 different stone patterns were sculpted in
clay. A negative rubber liner was then produced from
the clay molds that could be moved and rotated within
each panel to avoid repetition on the final panels. To
hide the normally vertical joints of the precast concrete
panel system, small, shallow connecting pieces were
grouted into place to span the joint space and simulate
the appearance of stone masonry work. The structure
is three bays wide, framed with 10 ft (3 m) wide double tees and precast concrete core walls. The core walls
serve as lateral support, with inverted tee beams and
columns picking up the center span’s load. Two fivestory wings are united by a central glazed rotunda that

2

(b)

Fig. 2.5.16(a), (b) & (c)
Starz Encore Headquarters, Englewood, Colorado; Architect: Barber Architecture;
Photos: (a & b) David Cornwell Photography, (c) The Weitz Company.

(a)

(c)

serves as the main visitor entry. The wings are basically
identical but are rotated 180 degrees from each other
so the building avoids redundancy with an alternating
convex/concave massing. The entablatures across the
front and back entries to the building consist of historically accurate, tapered precast concrete columns that
were cast vertically with horizontal joints to emulate
Roman/Tuscan columns. The columns align with the
belt course of the base panels (Fig. 2.5.16[b]). From initial design to full occupancy took only 22 months, with
the 308,000 ft2 (28,600 m2) office structure erected in
218 calendar days (Fig. 2.5.16[c]).

ARCHITECTURAL PRECAST CONCRETE

|

89

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.5.3 Design Considerations

The apartment complex in Fig. 2.5.17(a) was the first significant application of the precast concrete hybrid moment resistant frame (PHMRF) system. The PHMRF is
a lateral–force resisting building structural system using
precast concrete beams and columns joined together with
both mild steel reinforcement and post-tensioned highstrength steel strands. Compared to the common structural steel or cast-in-place concrete systems, this manner

(c)

(a)
Fig. 2.5.17(a), (b) & (c)
The Paramount, San Francisco, California;
Architect: Elkus Manfredi Architects, Ltd., Design Architect;
Kwan Hemmi Architects and Planners, Architect of Record;
Photos: (a & c) David Wakely Photography, (b) Pankow.

90

|

ARCHITECTURAL PRECAST CONCRETE

(b)

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.5.3 Design Considerations

of resisting seismic forces significantly reduces the potential
for damage to the key elements of a building’s frame. The
perimeter precast concrete moment frame serves as both
the lateral–force resisting system and the architectural façade (Fig. 2.5.17[b]). The architect was able to use the high
quality and flexibility of precast concrete to create an articulated and visually interesting façade design. Sandblasting,
reveals, and bullnoses all add to the complexity and interest
of the façade treatment (Fig. 2.5.17[c]). At 39 stories and
420 ft (128 m) high, the building was the tallest concrete
structure at the time it was constructed, in addition to being the tallest precast, prestressed concrete framed building
located in Seismic Zone 4. This building is the culmination
of the Precast Seismic Structural Systems (PRESSS) research
program that has been confirming the validity of advanced
precast concrete seismic concepts over the last decade.

2

and beams feature various levels of sandblasting and reveals. One hallmark of the project was the speed of construction, which is critical in a downtown area. Achieving
the completed look to the precast concrete structure took
only three-and-a-half months (Fig. 2.5.18[b]).
(a)

The seven-story mixed-use residential project in Fig.
2.5.18(a) features 225 lofts perched above ground floor
retail and restaurant space and three levels of above-grade
and subterranean parking. The total precast concrete
building was built with a precast concrete hybrid moment-frame system. A well controlled structural testing
program confirmed that this frame system demonstrated superior performance to a conventional cast-in-place
concrete frame. The building will be able to handle large
seismic drifts while experiencing minimal damage because
the structural system was designed with a post-tensioned,
self-righting mechanism. The system’s prestress design has
enough pre-compression to assure recentering of the building to its original position after the ground stops shaking
from a major seismic event. The precast concrete columns
(b)

Fig. 2.5.18(a), (b) & (c)
800J Plaza Lofts; Sacramento,
California; Architect: LPA
Sacramento Inc.;
Photo: (b) LPA Sacramento Inc.
(c)

ARCHITECTURAL PRECAST CONCRETE

|

91

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.6 Precast Concrete Panels Used As Shear Walls

2.6 P
 RECAST CONCRETE PANELS USED
AS SHEAR WALLS

(b)

In many structures, it is economical to take advantage
of the inherent strength and in-plane rigidity of exterior precast concrete wall panels by designing them to
serve as a part of the lateral-load resisting system. The
wall panels provide all or a portion of the lateral stability of a structure when combined with diaphragm
action of the floor construction. Walls taking in-plane
horizontal loads (lateral forces) from the effects of
wind or seismic forces are referred to as shear walls.
The shear walls may be, but do not need to be, bearing walls. Shear walls are used as the most common
and economical lateral–force resisting system and have
been used widely in buildings up to 30 stories tall, although more typically in low- to mid-rise structures.
Fig. 2.6.1(a) & (b)
Sarasota County Judicial Center, Sarasota, Florida;
Architect: BMK Architects Inc. and HLM Design, Joint Venture
Photos: Scott McDonald ©Hedrich Blessing.

(a)

A shear-wall system’s effectiveness is dependent
largely upon panel-to-panel connection design. A
significant advantage of jointed construction is in the
inherent ease of defining load paths through connections. As such, it is relatively easy to integrate a precast
concrete lateral–force resisting system’s performance
with that of the vertical loadbearing frame.
Shear walls are vertical members, which resist and
transfer lateral forces, in or parallel to the plane of the
wall, from the superstructure to the foundation. Thus,
shear walls act as vertical cantilever beams. Shear walls
are placed at appropriate locations within and around
the building perimeter according to the architectural and functional design requirements. The 1 ft (0.3
m) thick panels on the judicial center in Fig. 2.6.1(a)
measure 21 x 15 ft (6.4 x 4.6 m). They serve as shear
walls at the corners of the building (Fig. 2.6.1[b]).

92

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.6 Precast Concrete Panels Used As Shear Walls

2

Continuous steel plate connections were cast into the
corner panels to permit a welded connection at the
vertical corner joint.
Typically, a structure incorporates numerous walls, which
can be used to resist lateral forces, in both principal axes
of the building (Fig. 2.6.2). Connections can be designed
for specific directional resistances while maintaining flexibility, which may be beneficial. Because of the importance of shear walls in the behavior of the building, the
engineer of record must collaborate with the precast concrete engineer in the implementation of the lateral force
resisting system design. If the structure includes architectural precast concrete panels that could act as shear walls,
but are not intended to do so, the connections must be
designed so as not to attract unintended forces into the
panels. The portion of the total lateral forces that each
intended shear wall resists depends on the wall’s bending-and shear-resistance capacity, the participation of the
floor, and the characteristics of the foundation. For most
structures, lateral load distribution to the walls is based
on the properties of the walls (their relative stiffnesses).
The lateral forces the building must resist may be wind,
seismic or blast loads. The magnitude of these loads
varies according to a project’s purpose and geographic
location. Concrete panels have the inherent strength required to perform as shear walls. It is important that the
connections be designed to transfer lateral forces, and
also accommodate thermal movements and differential
deflections (or camber), as covered in Section 4.5.2. In
some cases, the ability to transfer lateral forces may be a
panel’s only structural purpose. But, it is more often combined with loadbearing or wall-supporting capabilities.
Shear walls are economical because walls already required by the building layout (such as exterior walls, interior
walls, or walls of the elevator, stairway, mechanical shafts,
or cores) can be designed as structural shear walls. Load
transfer from horizontal diaphragm to shear walls, or to
elevator walls, stairway cores, or mechanical shafts, can be
accomplished either via connections or by direct bearing.
Whenever possible, it is desirable to design shear walls as
gravity loadbearing panels. The increased dead load acting
on the panel is an inherent advantage to its function as a
shear wall because it increases the panel’s resistance to uplift and overturning forces created by lateral forces.
The effect of cumulative loads on connections between panels must be considered since these loads
become a significant factor in determining minimum

Fig. 2.6.2 Exterior shear walls.

panel dimensions. Shear walls in precast concrete buildings can be individual wall panels or wall panels that
are connected together to function as a single unit.
Connected panels greatly increase shear resistance capacity when compared to the same length of panels
acting independently as several narrower shear walls.
Connecting long lengths of wall panels together, however, can result in an undesirable build-up of volume
change forces. Hence, it is preferable to connect only as
many units as necessary to resist in-plane shear forces
and the associated overturning moment. Connecting as
few units as necessary near the mid-length of the wall
will minimize the volume change restraint forces.
In some structures, it may be desirable to provide shear
connections between non-loadbearing and loadbearing shear walls in order to increase the dead load resistance to moments caused by lateral loads. However, in
most cases, an exterior shear wall (or perimeter frame)
system provides more efficient and flexible floor plans
than an interior shear wall system because it may eliminate the need for a structural core.
Furthermore, exterior shear walls do not affect the
interior traffic flow or sight lines. The exterior walls can
be designed to provide the vertical strength and horizontal connections to allow the entire wall to function
as a single unit to mobilize dead load overturning resistance. In addition, they may eliminate the need for
exterior columns and beams (Fig. 2.6.2).

ARCHITECTURAL PRECAST CONCRETE

|

93

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.6 Precast Concrete Panels Used As Shear Walls / 2.7 Precast Concrete As Forms For Cast-in-place Concrete

element. Seismic and wind loads are resisted primarily by the building central core and partly by the ductile concrete exterior frame. Basically, floor slabs act as
diaphragms. Figure 2.7.1 illustrates the use of cast-inplace concrete to tie the walls, beams, and floor together. This can be an efficient system for providing
lateral resistance in precast concrete buildings.

Fig. 2.6.3 Interior shear walls.

In an interior shear wall system, the lateral forces are
not transferred directly to the foundation. Instead,
the wall panels distribute the lateral forces to floor
diaphragms, which, in turn, transfer them to a structural core or to the interior shear walls (Fig. 2.6.3).
Frequently, shear wall panels are connected vertically
and at the corners to form a structural tube that cantilevers from the foundation, making the panels more
efficient at resisting lateral loads.

2.7 P
 RECAST CONCRETE AS FORMS FOR
CAST-IN-PLACE CONCRETE
Architectural precast concrete units also may serve
as forms for cast-in-place concrete. This application is
especially suitable for combining architectural (surface
aesthetics) and structural functions in loadbearing façades (it avoids the problems of matching the surface
finish of the cast-in-place concrete to architectural precast concrete), or for improving ductility in locations of
high seismic risk by using wet cast connections with
high levels of reinforcement at the joints. Continuity
and ductility are achieved by casting in place the beams
and columns using precast concrete loadbearing panels as the exterior formwork.
The ductility of walls partially depends on reinforcement locations. Ductile behavior is improved significantly if the reinforcement is located at the ends of
the walls. This way, structurally inactive cladding can
be designed to become a major lateral load resisting

94

|

ARCHITECTURAL PRECAST CONCRETE

The use of precast concrete as forms can reduce construction time since all of the formwork required for a
structure can be manufactured in advance of concrete
placement. This permits greater flexibility and continuity in concrete placement activities. Delays in placing
the concrete due to the time required for preliminary
curing of concrete preceding form removal and re-erection of forms can be eliminated. The precast concrete
units may be erected quickly and the structure is complete when the cast-in-place concrete is placed and has
achieved its design strength. The need for temporary
outside forms is eliminated.
The architectural precast concrete form can be noncomposite and serve only to achieve a desired architectural effect after the cast-in-place concrete has
achieved design strength. This is accomplished by
providing compressible joints between abutting precast concrete panels, and neglecting (or eliminating)
bond at the interface of the precast and cast-in-place
concrete. The architectural precast concrete element is
then non-composite with the cast-in-place concrete.
Reinforcing steel extending from the precast concrete
into the cast-in-place concrete only needs to be of sufficient strength to support the formwork unit.
Realistic assumptions with regard to construction
techniques are required. It must be determined (or
specified) how and where the precast concrete panels
will be supported during concrete placement in order
to design the proper reinforcement within the panels. A mockup section may be necessary to test the
construction procedures before the project gets under
way. Such a mockup will also assist in refining panel
shape, size, finish, joint locations, and connections.
Concrete stay-in-place form panels should be erected
and temporarily braced to proper elevation and alignment in such a way that the tolerances specified for the
finished structure can be met. Temporary bracing for
the panels generally consists of adjustable pipe bracing
from panel to floor slab. Supports, braces, and form ties
must be stiff enough so that their elastic deformation
will not significantly affect the assumed distribution of

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.7 Precast Concrete As Forms For Cast-in-place Concrete

normally be the responsibility of the contactor. The designer should require that shop drawings be submitted
for review before the concrete is placed.

Fig. 2.7.1 Precast concrete as formwork.

(a)
Vertical section
•C
 ast-in-place portions
integrate with precast
to result in complete
lateral system.
• Wall panel and slab
shapes are planned
to create self-forming
elements.
• If precast concrete is
to be composite with
cast-in-place (C.I.P.)
concrete interior surface must be clean and
could be roughened
and/or have inserts for
ties bolts.

Horizontal section

• Inserts to facilitate
formwork may be
provided.
•R
 einforcing and inserts
must be thoroughly
planned to avoid
conflicts.
• Temporary shoring
and bracing may be
required.

Variations

(b)

2

Precast permanent formwork
(one-piece spandrels may
require support and restrict
placement of concrete)
C.I.P. Slab
Temporary formwork
Bond ties screwed into
cast-in inserts

the load from the fresh concrete. Form ties may be attached to embedded strap anchors or threaded inserts
provided in the panels for that purpose, or welded to
plates cast in the panels. Ties are then fastened in the
conventional manner with hardware on the outside of
the interior wood or steel forms. Column forms may
use column clamps or be wrapped with steel bands to
aid in resisting hydrostatic pressure. Care must be taken to protect corners of precast concrete units when
wrapping forms.
The designer should specify surface finish and desired
minimum thickness of precast concrete, but design
and layout of the forms and supporting systems will

In other cases, it may be desirable to establish interaction between the precast concrete form and the
cast-in-place concrete so that they act compositely in
the completed structure. It is then necessary to provide
shear transfer between the precast concrete and the
cast-in-place concrete. Shear transfer is accomplished
by bond and/or mechanical connections. The element
may then be treated as an integral unit for subsequent
applied loads, designed in accordance with the composite concrete section of the ACI 318 Building Code.
If the precast concrete element is arranged vertically
(such as a column or wall form) and not otherwise supported, the reinforcement that passes across the interface should be adequate to support the architectural
precast concrete unit. This reinforcement should be
adequate to restrain bowing in the precast concrete
element.
Deflection of stay-in-place precast concrete beam
forms, and warping of wall forms, may result from
differential shrinkage of precast and cast-in-place concrete, as well as from the dead load or lateral pressure of the cast-in-place concrete. Stay-in-place forms
should be designed to limit form deflection to 1/360 of
the unsupported height or length. Cambering of architectural precast concrete forms to compensate for
deflections is expensive and should be avoided. Where
the member is long enough to develop bonded strand,
pretensioning may be used in the precast concrete
form units.
Horizontal construction joints in the cast-in-place concrete should be made 3 in. (75 mm) below the top edge
of panels used as permanent forms rather than in line
with horizontal form joints. This reduces the possibility
of water leakage through the construction joints.
Design of composite flexural members using precast
concrete as forms requires locating form joints in areas
remote from points of high moment, any reinforcement in the precast concrete must be discontinued at
the form joint location. With the joint so located, the
shear at that section can usually be adequately resisted
by the reinforcement designed for this purpose and
contained in the cast-in-place concrete.
To determine the feasibility of architectural precast
concrete form units and the economies they may ef-

ARCHITECTURAL PRECAST CONCRETE

|

95

2

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS
2.7 Precast Concrete As Forms For Cast-in-place Concrete

fect, the following aspects of a structural concept
should be considered:
1. T he parts or elements of the structure that appear
to be most readily adapted to construction using
precast concrete forms.
2. T he types of form units best suited for the various
parts of the structure.
3. D
 esign and installation details for form units that
will fill their functional requirements with minimum
production and erection costs.
4. T he minimum size of structure or the minimum
number of form units for economical form unit
production.
5. R
 eduction or elimination of special form handling equipment not otherwise required for
construction.
6. S tructure modifications (details and dimensions) to
reduce number of odd shapes and sizes of form
units.

and is constructed of precast concrete panels that double as formwork for cast-in-place concrete. The precast
concrete kit-of-parts includes twin column panels 16 to
28 ft (4.9 to 8.5 m) high, a U-shaped back piece, and
9-ft-high (2.7 m) spandrels (Fig. 2.7.2[b] and [c]). The
hybrid columns have a sandblasted face. The loadbearing capacity of the integrated precast and cast-in-place
concrete frame resulted in a substantially stronger
wall, one that is structurally independent of the central
library core. The colonnade was used to evoke the classical language of traditional library architecture.
The 900,000 ft2 (84,000 m2) office complex in Fig.
2.7.3(a) and (b) consists of two terraced towers, nine
and ten stories high, that used precast concrete wall
panels as forms for the cast-in-place concrete. The wall
panels were designed with their edges serving as forms
for the columns and spandrel beams, thus integrating
the panels into a tube to resist lateral forces.
The exterior architectural expression of the medical
facility, characterized by large window openings and

The four-tiered elliptical freestanding colonnade in
Fig. 2.7.2(a) contains study areas and reading rooms,

Fig. 2.7.2(a), (b) & (c)
Library Square, Vancouver, BC, Canada;
Architect: Moshe Safdie and Associates;
Downs/Archambault Partners, Associate Architect;
Photos: (b & c) Downs /Archambault Partners.

(b)

(c)
(a)

96

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN CONCEPTS RELATED TO USAGE AND ECONOMICS

2.7 Precast Concrete As Forms For Cast-in-place Concrete

thin precast concrete spandrel and column profiles,
was established in the master plan phase of the project in Fig. 2.7.4(a). Thin spandrel and column profiles
seem to conflict with the functional requirement for a
structure that satisfies stringent vibration characteristics. However, both of these objectives were achieved
by using the 41/2-in.-thick (110 mm) precast concrete
spandrel beams and columns as part of the formwork
for the cast-in-place concrete (Fig. 2.7.4[b] and [c]).
This allowed the use of thinner precast concrete façade
elements and eliminated the space normally required
for precast concrete connections and the associated tolerances needed to satisfy the vibration criteria.

2

Fig. 2.7.3(a) & (b)
Marathon Plaza; San Francisco, California;
Architect: Whisler-Patri;
Photos: Pankow.

(c)
(a)

(b)
Fig. 2.7.4(a), (b) & (c)
University of Texas Southwestern Medical Center,
Phase IV
Dallas, Texas;
Architect: Omniplan Inc.;
Photos: Omniplan Inc.

ARCHITECTURAL PRECAST CONCRETE

|

97

3

Energen Building, Birmingham, Alabama;
Architect: Smallwood, Reynolds, Stewart, Stewart & Associates, Inc.; Photo: Gabriel Benzur.

CHAPTER THREE
SURFACE AESTHETICS
3.1 G
 ENERAL
Many facets in the design of architectural precast
concrete are of vital importance to the architect. Two
significant design considerations described in Chapter
2 were total wall analysis and repetition and the master
mold concept. Chapter 3 discusses the surface aesthetics of precast concrete panels, which require decisions
by the architect on considerations such as color, form
and texture, and weathering. Because of the versatility
of the material, the architectural focus can vary greatly
from project to project, changing the relative importance of each of these facets to the design.
Proper selections of color, form, and texture for a
building’s precast concrete exterior is critical to creating a successful aesthetic appearance. The decisions
depend not only on cost, delivery schedule, and client
satisfaction but on the local and regional context as
well. The desired colors and textures can be achieved
by varying aggregate selection, matrix color, finishing
processes, and depth of exposure of the aggregate.
The proper use of samples and mockups can ensure
the project’s success.

samples and recognizing uniformity requirements should
be considered, and the designer should focus on selecting shapes, sizes, colors, textures, and finishes for the
samples well in advance of finalizing the bid documents.
The building’s appearance results from the architect’s
use of light, shadow, texture, and color. Color and,
consequently, color tone represent relative values. They
are affected by light and shadow, intensity, time of
day, and nearby colors. Thus, color selection should be
made in lighting that replicates the light and shadows
of the site’s natural daylight.
The architect should give sufficient details or descriptions on the contract drawings to indicate clearly the
extent of all exposed surfaces of the units and their
respective finishes. This is particularly important for
returns and interior finishes. The location and dimensions of reveals should also be shown. All of these
items should then be shown on the shop drawings.

3.2 UNIFORMITY AND DEVELOPMENT
OF SAMPLES

However, because of different material sources and
manufacturing techniques, the guide’s photographic
samples and the final product may not be an exact
match. Samples must be made to ensure that the desired colors and textures are satisfactorily matched.
Samples for architectural precast concrete are custom
produced to translate the architect’s specific design
concept into a standard for realistic and economic production requirements.

Because acceptable color uniformity and shading intensity are evaluated visually, they are generally a matter of an individual’s subjective judgment and interpretation. Acceptable variations in color, texture, and
uniformity should be determined at the time the sample, mockup, or initial production units are approved.
Accordingly, it is beyond the scope of this Manual to
establish precise or definitive rules for product acceptability on the basis of appearance. However, a suitable criteria for acceptability requires that the finished
concrete surface should have a pleasing appearance
with minimal color and texture variations from the approved samples. The finished face surface should show
no obvious imperfections other than minimal color and
texture variations from the approved samples or evidence of repairs when viewed in typical lighting with
the unaided eye at a 20 ft (6.1 m) viewing distance.
Appearance of the surface also should not be evaluated when light is illuminating the surface from an extreme angle, as this tends to accentuate minor surface
irregularities (see Section 3.5.17).

In the schematic design stage, a schedule for creating

The major factors affecting uniformity of architectural

Precast concrete allows architects to be innovative
and create designs that cannot be accomplished with
other materials. It provides the freedom and flexibility
of shaping concrete into structure and architecture.
The Architectural Precast Concrete–Color and Texture
Selection Guide, published by PCI, helps architects define and achieve their aspirations. The guide’s photographs serve as a visual reference for initial selection
of color, texture, and finish and should be followed by
producing samples at a precaster’s plant to aid in the
final selection of color and texture.

ARCHITECTURAL PRECAST CONCRETE

|

99

3

SURFACE AESTHETICS

3.2 Uniformity and Development of Samples / 3.2.1 Uniformity

precast concrete units are described in Section 3.2.1.
These should be recognized through all stages necessary to prepare, assess, and approve samples.
Samples of architectural precast concrete are intended
to represent the materials and finish used. The concrete’s
color or appearance likely will vary during production,
so samples showing that expected range should be required. Product appearance is influenced by factors such
as quality, complexity of the casting, and physical mass,
as well as the natural characteristics of the concrete ingredients. In short, a single 12 × 12 in. (300 × 300 mm)
sample may not accurately represent a production casting, so larger samples should be used.

3.2.1. U
 niformity
Concrete contains natural materials, and it is these
materials’ inherent beauty that is most often expressed
in architectural concrete. The limitations of these natural materials with respect to uniformity must be considered, and the requirements for uniformity of the precast
concrete product must be set within these limitations.
Some color difference between nominally identical
precast concrete units is inevitable, but color variation,
between and within panels, should be kept within an
agreed range. Therefore, it is important, at the sample
stage, to reconcile the expectations of the owner and
architect with the practical limits of color uniformity.
Some designers prefer to see color variation akin to
timber and natural stone, while others desire the consistency and uniformity of paint. Where uniformity is
essential, the precaster can provide significant input in
balancing colors, textures, and shapes to achieve this
uniformity.
Color control is, thus, about ensuring that panels or
other precast concrete elements for a project have an
acceptable tonal range.
Uniformity of color and texture requires the precaster
to manage a complex set of variables, including raw
materials, mixture proportions, mixing, casting and
consolidation, curing, finishing, and weathering. When
fabrication continues over extended periods, color can
vary because of the changes in the physical characteristics of cements, coarse aggregates, and sand, even
though they may be from the same sources.
The color of a concrete is dependent on, among other
factors, the cement and other materials used. Variation
in the color can occur from day to day in the product

100

|

ARCHITECTURAL PRECAST CONCRETE

from a single cement plant, and color differences are
to be expected among cements obtained from different plants. Cement color reflects chemical composition
and processing conditions. Usually, cement colors vary
from white to shades of gray and brown. Greater color
uniformity results can be expected when using white
cement than when using gray cement.
The type and brand of cement must also remain consistent. Changing from Type I to Type III portland cement within one job will cause color variations because
Type III portland cement is a finer grind of cement than
Type I. Even though the color changes of the cement
would be minimal, it is recommended that types of cement not be changed.
Because the largest portion of a concrete mixture is
aggregate, the color or gradation of aggregate can influence the color of concrete. A substantial change in
aggregate color can make a noticeable difference in
surface color, especially if an exposed finish is specified. Therefore, the precaster should stockpile, either
at the plant or quarry, the fine and coarse aggregates
for each type of exposed finishs.
Coarse aggregates should be reasonably uniform in
color. A mixture can have more than one aggregate
type to get the desired color. Light and dark coarse
aggregates require care in blending so that color uniformity is achieved within a single unit. Choosing aggregates with a small color difference between the
light and dark aggregate will enhance uniformity. The
architect should specify that the matrix’s color or tone
match that of the coarse aggregate so that variations
in the depth of exposure and concentration of aggregate will not be as noticeable. Panels containing aggregates and matrices of contrasting colors will appear
less uniform. Also, as the size of the coarse aggregate
increases, less matrix is seen.
The fineness modulus (FM) of the fine aggregates and
the content of fines (particles passing a #200 [75 µm]
mesh sieve) can have a significant impact on final appearance. Units made with a higher content of fines will
be lighter colored due to the increased surface area of
fine particles and their light-scattering characteristics.
Elimination of fines, or keeping them to a minimum,
will help to prevent color variations.
Chemical admixtures and pigments affect final color.
They need to be added in the same amounts and in
the same sequence throughout the job to avoid color
variations.

SURFACE AESTHETICS

3.2.1 Uniformity

Each factor discussed here affects color consistency,
but daily variations in moisture content are probably
the single most common cause of color consistency
problems. A change in the water-cement ratio can result in color inconsistency from batch to batch.
The water-cement ratio in a concrete batch is affected
by the moisture content of the raw materials, primarily
the sand, and the amount of mixing water. Automatic
moisture control of the sand and adjustment of the
mixing water volume for every batch help to minimize
such color fluctuations.
The mixing time required to achieve complete dispersion of all materials varies from plant to plant depending on the type of mixer and the aggregates used. If
pigmented concrete is not mixed long enough, the
color is less intense. Also, if the concrete batching sequence is varied, color uniformity will be affected.
The color of precast concrete can vary between adjacent elements due to daily variations in the curing
conditions for the concrete. The concrete and mold
temperatures should remain as consistent as possible
throughout the job to minimize color variations.
If a sample is stored indoors, its color will vary from
a panel stored outdoors. A panel stored outdoors and
exposed to precipitation is cured differently than the
controlled enviornment of the sample. It is difficult to
exclude the influence of the climatic changes on color
over a year if the precast concrete units are placed in
storage for long periods of time, as may be dictated
by contractural conditions or by operations at the construction site beyond the control of the precaster.
The last production process that affects panel aesthetics and needs to be controlled is the finishing. A
smooth-off-the-form finish is extremely difficult to produce consistently. Any type of finish that has some degree of aggregate exposure will appear more uniform
than a smooth finish because the natural variations in
the aggregates will camouflage subtle differences in
the texture and color of the concrete. The degree of
uniformity normally improves with an increased depth
of exposure. Some variation is to be expected in color
and texture, even after finishing. Assessment of color
uniformity of the panels prior to finishing offers little
information. Dividing large surface areas into smaller
ones with reveals or rustications also helps to lessen
any variation in texture that might be visible.

3

Many finishes cannot be achieved with equal visual quality on all faces of the unit because of several
factors, such as mixture proportions, variable depths
(pressures) of concrete, and differences in consolidation techniques, particularly in the case of intricate
shapes with complex flow of concrete.
During consolidation, the effect of gravity forces
the larger aggregates to the bottom and the smaller
aggregates, plus the sand and cement, upwards.
Consequently, the down-face in the mold will nearly
always be the most uniform and dense surface of the
unit. The final orientation of aggregates may also result in differences in exposure between the down
face and the returns in exposed-aggregate surfaces.
Emphasis should be placed on choosing suitable concrete mixtures with aggregates that are reasonably
spherical or cubical in shape to minimize differences.
For large returns, or situations where it is necessary to
minimize variations in appearance, concrete mixtures
should be selected where the aggregate gradation can
be uniformly controlled and preferably fully graded.
Exposures should be medium to deep, and color differences between the ingredients of the mixture should
be minimal.
The color of any concrete product can be expected to
change to some degree over time. Atmospheric pollution and any accumulated grime or soot will darken the
surface. These effects can be controlled by producing
well-detailed precast concrete units with high-quality
concrete. Just like all material surfaces left in the open,
precast concrete occasionally must be cleaned to remove pollutants and restore color. Efflorescence may
occur randomly on the product surface during its first
several years of exposure, which can cause it to look
faded or lighter in color if not cleaned off. After years of
exposure, the cement paste may erode from the surface
depending on environmental conditions, such as acid
rain. This will expose more fine aggregate and shift the
color of the concrete to the color of the aggregate.
The sample’s appearance should be assessed during
both wet and dry weather. White concrete usually produces less of a difference in tone between wet and
dry panels. In climates with intermittent dry and wet
conditions, drying-out periods may produce temporary
mottled appearances in all-gray cement façades, particularly on fine-textured surfaces. On the other hand,
dirt (or weathering) normally will be less objectionable
in gray panels.

ARCHITECTURAL PRECAST CONCRETE

|

101

3

SURFACE AESTHETICS

3.2.1 Uniformity / 3.2.2 Development of Samples

Some factors are outside the precaster’s control:
• Changes in cement color. This is more likely to be
associated with gray cements than with off-white
or white because the latter are manufactured to
very close color tolerances.
• Variations in curing as a result of changes in ambient temperature and humidity.
• Variations between horizontally and vertically cast
units.
Although material and production factors may cause
differences in color or texture, lack of uniformity will
be minimized if the recommendations of this section
are followed. These include creating pre-bid samples to
establish the general color and texture for the project,
producing approval samples after the contract award
to evaluate the same mixture under sample production
conditions, producing 4 × 4 ft (1.2 × 1.2 m) sample
panels to show the range of anticipated color and texture, and viewing initial production panels to see the
final outcome of the process based on bulk ordering of
currently quarried materials and full concrete batches.

3.2.2 D
 evelopment of Samples
For the architect to develop and select the color and
texture for architectural precast concrete at the conceptual stage requires a combination of art and skill.

Fig. 3.2.1

102

|

ARCHITECTURAL PRECAST CONCRETE

The same is equally true of the precaster, which must
translate these requirements into workable concrete
mixtures and the proper finishing techniques. There
are numerous choices in textures and colors due to
the great range of coarse aggregates, sands, cements,
and pigments, combined with a variety of finishing
processes.
Achieving the desired textures and colors with feasible production techniques is a process that requires
the precaster to produce samples that satisfy the architect’s design concepts. This can be accomplished
by producing a few samples, or it may require a series of samples and considerable investigation of corresponding production and finishing techniques or by
reference to the PCI Color and Texture Selection Guide.
Figure 3.2.1 shows various samples to assist in selecting architectural finishes for shadowing and color. The
selected samples should be available for inspection and
examination by prospective bidders.
The importance of this process is not always recognized by architects or precasters. To ensure success, all
research and development, including mixture proportioning, should be completed prior to formal bidding
so that all precasters are estimating similar materials
and finishes. It is also recommended that all precasters
approved for a particular project develop samples for
approval as a prerequisite for bidding.

SURFACE AESTHETICS

3.2.2 Development of Samples

Polished

Sandblasted – Light

Sandblasted – Light

Retarded – Medium

Retarded – Medium

Acid Etched – Light

Polished

3

Acid Etched – Light

Fig. 3.2.2 Appearance variations achieved with different finishes on the same concrete mixture.

At this stage of the procedure, the development of
samples may involve considerable expense in research
and investigation on the part of the precaster. The architect can aid in sample development by visiting precasting plants that have sample selections on hand to
assist in selecting limits for the desired finish. Because
the architect is responsible for the final decision, design judgment should be supplemented with an assessment of the operating procedures and technical
personnel from all plants likely to bid on the project.
Watching plant operations and talking with plant personnel also help the architect obtain an understanding
of production considerations.
Samples should be at least 12 × 12 in. (300 × 300
mm) to provide information on face mixture proportions
(color tone) and finishes (texture) for the architect’s initial aesthetic evaluation (Fig. 3.2.2). Larger samples are
recommended, but they may be difficult to handle. The
size of the samples should relate to the maximum size
of aggregate to be used to allow for realistic placement
of the concrete and accurate expression of detail.

• Degree of surface texture or depth of aggregate
exposure.
Color selection should be made under lighting conditions similar to those under which the precast concrete
will be used, such as the strong direct light or shadows
of natural daylight.
Both designers and owners should remember that
selection of a precast concrete sample represents only
the first step in the development of the actual production of that element. It should not be considered a final decision. Completing the sample process remains
extremely important and develops communication
among all parties.

• Matrix color (cement, pigment, and sand color).

Some precasters have small samples in stock to show
the colors, finishes, and textures used on previous projects (Fig. 3.2.3). Previous work of a similar nature can
serve as a useful visual standard and highlight potential
concerns. Even though an architect has seen the selected aggregates used with a similar finish in existing
precast concrete units, it is important to develop specific project samples. These samples must reflect the
relationship between materials, finishes, shapes, and
casting techniques, such as mold types, orientation
of exposed surfaces during casting, and consolidation
procedures.

• Coarse aggregate type and source (where aggregate exposure is planned).

It is recommended that reference samples be used to
determine product characteristics and quality, rather than

From the 12 in. (300 mm) square samples, a preliminary evaluation should be made on the following
issues:

ARCHITECTURAL PRECAST CONCRETE

|

103

3

SURFACE AESTHETICS

3.2 Development of Samples / 3.2.4 Production Approval Samples

1. Sufficient time must be allowed for the bidder to submit samples or information
for approval. Time also must
be provided to allow the approvals to be conveyed to the
precaster in writing so they
can estimate and submit an
accurate bid.
2. Any pre-bid submittal should
be treated in confidence,
and the individual producer’s
solutions and/or techniques
should be protected both
before and after bidding.
All submitted samples should
be clearly identified with the precaster’s name, date produced,
Fig. 3.2.3
identifying code number, and
Samples showing color, finish and texture at precast concrete plant.
name of project for which it was
submitted. If the precast concrete
units are to have an exposed interior finish, samples
explicit specifications by the architect that might prohibit
should also be provided for this purpose.
the precasters from using a material or process that offers
the best solution for producing the desired results.
If the characteristics of submitted pre-bid samples

3.2.3 P
 re-Bid Samples
Individual plant preferences, differences in sources
of supply, or different techniques developed in various
plants serving the same area mean that not all precasters will be able to obtain an exact match of the
selected sample(s). Many architects select and approve
samples prior to bid closing. Then the approved precasters’ names and corresponding sample code numbers are published in an addendum or the approval list
is given in writing to the general contractors.
This practice may result in slight variations in color,
aggregate, or texture but not necessarily in the quality supplied by different bidders. The individual precaster, within specification limits, selects the materials
and employs the placing and finishing techniques best
suited to its plant operations. By making approval of
pre-bid samples a prerequisite for bidding, the architect and client are protected by requiring equivalent
optimum quality from all precasters. All involved then
know the result to be achieved in color and texture of
the finish. When making pre-bid approval of samples
part of the specifications, the architect should adhere
to the following requirements:

104

|

ARCHITECTURAL PRECAST CONCRETE

deviate from the project specifications, the precaster
must make this clear when submitting the samples and
other required information for approval. For proper
evaluation and approval of the samples, the precaster
should state the reasons for any deviations. These reasons might include the precaster’s concern over controlling variation in either color or texture within specified limits. In regard to adequacy of specified materials,
concerns about satisfying all conditions of the specifications must be based on practical plant production
requirements and the performance or weathering of
the product in its final location.

3.2.4 Production Approval Samples
After award of the contract but before producing any
units, the precast concrete manufacturer should prepare and submit for approval a representative sample or
samples of the required color and texture. This doesn’t
need to be done if the pre-bid samples prepared by the
plant were the basis for the specifications or the prebid approval method was used. Samples should be at
least 12 × 12 in. (300 × 300 mm).
Although 12 in. (300 mm) square samples provide
valuable information on texture and color tone for the

SURFACE AESTHETICS

3.2.4 Production Approval Samples

3

Fig. 3.2.4(a), (b) & (c)
Mockups to assist in finish selection.

(a)

architect’s initial aesthetic evaluation, these small samples are unlikely to give a true picture of the possible
variations of finish over a large area, demonstrate normal surface blemishes, or show the effects of the natural day-to-day variations of aggregates and cement.
Once the small samples are within an acceptable
range, larger samples should be made to confirm
that the mixture proportions, vibration, and finishing

(b)

techniques necessary to make production-sized pieces could duplicate the aesthetic qualities of the small
sample pieces.
For non-planar, curved, or other complex shapes, a
flat-cast sample may not represent the anticipated appearance of the final product. Sample shapes should
be selected to offer a reasonable comparison to the
precast concrete units they represent. Also, the size
of the samples should reflect the relationship among
finishes, shapes, and casting and consolidation techniques. These techniques include mold types, thickness
of concrete section, orientation of exposed surfaces
during casting, and consolidation procedures. If the
precast concrete units have an exposed interior finish,
samples of the finish, color, and texture should also be
shown for the back surface.
Any changes in material sources or in mixture proportions to facilitate production require new reference
samples and approval review. Samples showing the expected range of variations should be supplied if specified or if the color or appearance of the cement or the
aggregate is likely to vary significantly.

(c)

Figures 3.2.4(a), (b), and (c) show that a 12 in. (300
mm) square sample with a 2 in. (50 mm) thickness may
bear little relationship to the appearance and physical characteristics of a production panel. Differences

ARCHITECTURAL PRECAST CONCRETE

|

105

3

SURFACE AESTHETICS
3.2.4 Production Approval Samples

(a)

(b)

(c)

(d)

Fig. 3.2.5(a – d) Range samples.

in mass, density, and curing rate between the sample
and the production panel may make direct comparison
difficult. This is particularly true for insulated sandwich
panels.
Mockups that include all project mixtures and finishes, as well as all major details and reveals, can be used
to evaluate the production methods and the finished
product. For example, if return elements are to be cast
with a major panel section, the samples should have
returns cast with them to represent how the finish will
be accomplished on such sections. During review, features and problems unique to the project should be
discussed. General factors affecting any color and finish variations should also be discussed.
Range samples should be produced: (1) when required by specification, (2) past experience of the plant
with a mixture or finish, and (3) for large projects with
multiple approving entities with little apparent precast
concrete experience. At least three range samples of

Fig. 3.2.6 Production panel.

106

|

ARCHITECTURAL PRECAST CONCRETE

one size (full scale, but not necessarily full size) should
be produced to demonstrate actual planned production conditions. These should establish the range of acceptability for color and texture variations; uniformity
of returns; frequency; size and uniformity of air-void
distribution; surface blemishes; and overall appearance. During the range sample review, the precaster
will ask the architect or representative to inspect and
approve (sign and date) the range panels.
Samples or mockups should be viewed at a distance
consistent with their position on the structure but no
less than 20 ft (6.1 m). Overlooking this procedure can
result in demands for shapes, textures, and drafts that
are not only expensive but might not add value to the
final project. Approved 12 × 12 in. (300 × 300 mm)
samples should also be compared to the mockup to
ensure original intent has not been lost. During the
mockup review, the precaster will ask the architect or
representative to inspect and approve (sign and date)
mockup panel(s). Approved mockup panel(s) supersede the previously approved 12 × 12 in. (300 × 300
mm) samples. For panels to be properly evaluated, they
should be stored under conditions similar to those of
the actual work. Panels will be darker when damp than
when dry.
The approved range samples or mockup panel(s)
should be stored outdoors and positioned to allow
comparison with production units. They also should be
stored adjacent to each other to ensure similar lighting (sun and shade) for daily comparisons of finish and
exposure. Figures 3.2.5(a) to (d) show the acceptable
range of concrete samples made with 3/8 in. (10 mm)
aggregate with retarded (left side) and acid-etched
(right side) finishes. Only two of the retarded finishes
were within the acceptable range on the three acidetched samples. Therefore, one additional sample was

SURFACE AESTHETICS

3.2.4 Production Approval Samples

made to obtain an acceptable range for the retarded
finishes. Figure 3.2.6 shows the production panel made
after examining the samples shown in Fig. 3.2.5.

3

(a)

The acceptability of repair techniques for chips, spalls,
or other surface blemishes can also be established on
these samples. The face of each sample should contain at least two areas of approved size and shape that

Fig. 3.2.7(a), (b) & (c) Mockups.

(b)

have been chipped out and then patched and repaired.
The color, texture, and appearance of patched areas
should match that of the adjacent surface (see Section
3.5.17). In evaluating this experimentally repaired area,
the repair should be aged to give a true indication of
its color. Repairs to the mockup should be at least one
month old before acceptability is judged. Perfecting a
repair procedure can save both time and money in the
final outcome of the project.
If the project’s size warrants, the architect and owner
should authorize the expenditure for mockups, either
of a full-scale portion of a panel or the entire typical
unit. These may be several modules wide by one or
two stories high. Investing in such mockups removes
uncertainties held by both the architect and owner
and may lead to modifications that improve the appearance and possibly reduce the overall project cost
(Fig. 3.2.7[a] to [c]). The mockups allow the precaster
and designer to explore a series of options for particularly challenging details prior to full-scale production.
Larger samples require considerable time to produce,
and they should not be specified unless sufficient lead
time exists. If mockups are implemented in a timely
manner, cost and schedule implications associated with
revisions to the design may be avoided and measures
adopted promptly to rectify problems, if any. Also, it

(c)

may be desirable to separate the mockup costs from
the base bid so the cost can be evaluated separately.
The ability to satisfy building envelope performance
characteristics depends on the attention to detailing
at the interface of dissimilar materials. Because each
trade usually terminates its responsibility at the outer
boundary of the material, critical information can be
missed that affects the overall project. The mockup is
an ideal mechanism for coordination of all trades with
abutting materials.

ARCHITECTURAL PRECAST CONCRETE

|

107

3

SURFACE AESTHETICS

3.2.4 Production Approval Samples / 3.2.5 Assessment of Samples

Mockups can evaluate the following factors:
1. Range of acceptable appearance of color, texture,
and details on the exposed face and uniformity of
returns.
2. Orientation of casting (necessity of sequential casting).
3. Erection and bracing techniques.
4. Connection details.
5. Colors and finishes of adjacent materials (for example, window frames, glass, and sealants).
6. Dimensional accuracy of the precast concrete work
and the constructibility of the specified tolerances.
7. Acceptability of the precast concrete unit’s inside
surface finish (where exposed).
8. Suitability of the selected sealers, if applicable.
9. Weathering patterns or rain runoff on a typical section of the precast concrete façade.
Mockups should be produced using standard production equipment and techniques. Important variables
that should be controlled as close to actual cast conditions include retarder coverage rate and method of
application, if used, mixture design and slump, admixtures, temperature of fresh and cured concrete, vibration, piece thickness, age at which finishing operations
are performed, and method of cleaning. This is especially important with light etches that are particularly
affected by changing conditions. Special details, such
as reveal patterns and intersections, corner joinery, drip
sections, patterns, colors, and textures, should be demonstrated in the mockup units for approval. Changes
in aggregate orientation, color tone, and texture can
easily be noted on full-scale mockup panels.
The mockup sample also can demonstrate the more
detailed conditions that may be encountered in the
project, such as recesses, reveals, outside/inside corners, multiple finishes, textures, and veneers. Mockup
panels should contain all expected cast-in inserts, reinforcement, and plates.
When the mockups are manufactured and erected,
all interested parties should be present and ready to
discuss the approval for production of the panels. If
changes are desired, all information should be recorded. Depending on the changes, production should
not begin until the changes have been made and the
mockups are approved.
Where mockups are not used, the architect and/or

108

|

ARCHITECTURAL PRECAST CONCRETE

owner should visit the precast concrete plant and approve (sign and date) the initial production units. This
approval should precede a release for production to
avoid potential controversies later. However, delays in
visiting plants for approvals will upset normal operations and the job schedule. The contract documents
should state clearly how long the production units or
the mockup structure will be kept in the plant or at the
jobsite for comparison purposes. If specifications require mockups to be kept at the project site, sufficient
additional samples should be maintained for quality
control at the plant.
The contract documents also should permit the approved full-sized mockup units to be used in the job
installation in the late stages of construction. The units
should remain identifiable even on the structure until
final project acceptance. The panels should be erected
adjacent to each other to allow continued comparison,
if necessary.

3.2.5 Assessment of Samples
If 12 in. (300 mm) square samples are used to select
the aggregate color, the architect must remember that
the general appearance of large areas of a building wall
tend to be lighter than the samples. For example, exposed gray granite (salt and pepper) may look good on
a small sample, but frequently comes out “mottled” in
an actual panel if the coarse aggregate is small. If the
predominant part of the granite is white, the mottling
will be made worse with a gray matrix and vice versa.
The finish may be made more acceptable if the face
is sandblasted because of the resulting dulling of the
colors, but it is still better to increase the maximum aggregate size to eliminate visual merging of the colors.
Mockups are best assessed effectively when mounted
in their final orientation. Samples viewed from a distance
of a few feet will reveal details that will not be noticed
on a building when viewed from 50 to 100 ft (15 to 31
m). Details should be appraised from a distance typical for viewing the installed panel. Overlooking this may
lead to demands for shapes, textures, and drafts that
are not only expensive but may not even be identifiable
in the finished building.
Another good example is the fluted panel. When
viewed from a distance, the ribs should be reasonably
deep to read. They should also have a draft related
to the depth and spacing of ribs in order to facilitate

SURFACE AESTHETICS

3.2.5 Assessment of Samples / 3.2.6 Assessment of Concrete Mixtures

stripping without damage (see Section 3.3.2). Some
precasters advocating increased draft have shown
that draft does not detract from appearance by making panel samples with different drafts and by having
them evaluated by the architect from a distance typical
for viewing the installed panel.
The architect should observe the samples under the
climatic conditions to which precast concrete units will
be exposed, such as direct sun, rain, or shadows. The
samples should also be judged in relation to adjacent
buildings.
There is rarely enough time to allow weathering of
samples over an adequately long period, but it is particularly important where a project with precast concrete is contemplated for production in stages. The architect is advised to limit the choice of aggregates and
finishes in such projects to those that are in common
use and are easy to duplicate in later stages. To counter
weathering effects, cleaning of an earlier stage upon
completion of the next one will often provide a reasonable match.
To obtain a reasonable appearance uniformity, a balance may have to be struck between configuration of
the precast concrete unit and the choice of a concrete
mixture. Returns in some finishes will not appear exactly like the front face (down face) due to casting
techniques and aggregate shapes (see Section 3.3.7).
This should be recognized and accepted within certain limits because it may well influence the architect’s
choice of shape, materials, and finishes.
Difficult mixtures and finishes with respect to uniformity may be appropriate and economical in flat panels
cast face-down and without any appreciable return,
but not in highly sculptured panels.
The architect should look at the many existing precast
concrete applications and also recognize that added
variations and new design concepts are possible.

3.2.6 A
 ssessment Of Concrete Mixtures
The architect should specify the parameters of concrete performance requirements, but the actual design
of the concrete mixture should be left to the precaster.
So that the architect may appraise the appearance
and the expected performance of precast concrete
units using a specific mixture, information should be

3

obtained about the mixture to assess its anticipated
performance and appearance. Such an assessment
should be part of the pre-bid sample procedure described in Section 3.2.3.
This section discusses concrete characteristics to help
the designer specify the proper concrete requirements
and evaluate the mixtures proposed for the project.
All concrete mixtures should be developed using the
brand and type of cement, the type and gradation of
aggregates, and the type of admixtures proposed for
use in production mixtures. If at any time these variables are changed, the mixture should be reevaluated.
This reevaluation may include one or more of the following concrete properties: (1) color, surface texture,
or aggregate exposure; (2) air content or durability; or
(3) strength (selected tests at appropriate ages).
Face and Backup Mixtures. The use of a separate concrete face mixture and a subsequent backup
concrete or the use of a uniform concrete mixture
throughout a unit depends on the practice of the particular plant or the size and shape of the unit and the
type and extent of finish being produced as well as the
setback of the windows. For reasons of economy, face
mixtures are generally used, except in units of complex
shapes and deep, narrow sections or returns where the
procedures for separating the face and backup mixtures become too cumbersome. The choice should be
left to the precaster.
The face mixtures contain special decorative aggregates, often in combination with white portland
cement and pigment and are specially designed to
achieve the desired surface appearance. Backup mixtures are composed of more conventional aggregates
and gray cement and are used to reduce material costs
in large units that have a decorative face mixture. If a
backup mixture is not used, costs will be higher unless
economical aggregates are used. However, a face mixture will be used for the full thickness when the material savings do not warrant the added costs of working
with two mixtures.
Where a precast concrete unit is manufactured with
an architectural concrete face mixture and a structural
concrete backup mixture, these mixtures should have
reasonably similar shrinkage and thermal coefficient of
expansion characteristics in order to avoid possible undue bowing or warping. Consequently, these two mixtures should have similar water-cement and cementaggregate ratios.

ARCHITECTURAL PRECAST CONCRETE

|

109

3

SURFACE AESTHETICS
3.2.6 Assessment of Concrete Mixtures

The combination of a normalweight face mixture and
a backup mixture with lightweight aggregates may
increase the possibility of bowing or warping. Before
accepting such a combination of mixtures, sample
units which are produced, cured, and stored under anticipated production conditions are often desirable to
verify satisfactory performance.
If a separate face mixture is to be used, a minimum
thickness should be determined. The thickness of a face
mixture after consolidation should be at least 1 in. (25
mm) or a minimum of 1.5 times the maximum size of the
aggregate; whichever is the larger. If larger aggregates
are hand laid in the mold, these dimensions should apply to the concrete mixture used as the matrix.
The 1 in. (25 mm) dimension is chosen because the
consolidated face mixture is often used to support
the reinforcing steel cage and thus provide the proper
concrete cover over the reinforcement. For units not
exposed to weather or for face mixtures applied faceup (seeded), this dimension may be reduced provided
the backup mixture does not bleed through the face
mixture.
A concrete design strength should be determined
by the design team based on in-service requirements,
not forgetting production and erection considerations.
Because precasting involves stripping of units from the
mold at an early age, rapid strength development is
of prime importance. Transportation and erection involves the next strength requirement to which precast
concrete units are exposed. The precaster should establish minimum stripping and transportation strength
requirements. These strength levels will depend on the
shape of the unit, handling, shipping, and erection
techniques, and will normally result in a high 28-day
strength. A 5000 psi (34.5 MPa) compressive strength
at 28 days normally satisfies production requirements
and also ensures proper durability.
In cases where a 5000 psi (34.5 MPa) strength of the
face mixture is not structurally necessary, or is difficult to
attain due to special cements or aggregates, the architect may still achieve sufficient durability and weathering
qualities by stating proper air-entraining and absorption
limits at a strength level as low as 4000 psi (26.7 MPa).
The strength of face and backup concrete is usually
determined by using 6 × 12 in. (150 × 300 mm) or 4 ×
8 in. (100 × 200 mm) standard cylinders. If fabrication
of cylinders is impractical, 4 in. (100 mm) cubes may be
used. The measured cube strength should be reduced

110

|

ARCHITECTURAL PRECAST CONCRETE

20% unless strength correlation tests to 6 × 12 in.
(150 × 300 mm) cylinders have been made to obtain
an estimate of cylinder strength. It may be impractical
to prepare a standard test cylinder, for example, in the
case of a face mixture containing a high percentage of
coarse aggregate. The 4 in. (100 mm) cube will provide an adequate size for practically all face mixtures.
Such cubes may be prepared as individual specimens,
or they may be sawed from 4 in. (100 mm) thick slabs.
The slabs may be more convenient and are probably
more representative of the final product.
In assessing the strength of concrete, statistical probabilities should be considered. Many variables can
influence the strength of concrete even under close
control. The strength level of the concrete should be
considered satisfactory if the average of each set of
any three consecutive cylinder strength tests equals or
exceeds the specified strength and no individual test
falls below the specified value by more than 500 psi
(3.4 MPa). Alternatively, compressive strength results
from a predetermined number of consecutive tests
may be processed statistically and the standard deviation established. This approach will measure the
overall uniformity in performance. See also ACI 214,
Recommended Practice for Evaluation of Compression
Test Results of Field Concrete.
It is advisable to specify air-entraining requirements
for face mixtures in precast concrete units exposed to
freeze and thaw cycles in the presence of moisture. An
air entrainment of 4% is normally desirable. Taking into
consideration the many special consolidation techniques
used for placing face mixtures, a fairly liberal variation
of this percentage, such as +2% and -1% should be
allowed. The amount of air specified as 4% is thus acceptable if it is measured between 3% and 6%.
Because precast concrete units are generally erected
in an above-grade vertical position, which is a moderate environment, air contents as low as 3% to 5%
appear to provide the required durability. Low levels of
air entrainment are preferred because the compressive
strength of concrete is reduced by approximately 5%
for each 1% of entrained air (when the water-cement
ratio is held constant). Strength reductions due to air
entrainment tend to be greater in mixtures containing
more than 550 lb (250 kg) of cement per cubic yard.
Because most architectural precast concrete mixtures
contain a high cement factor, relatively high reductions
in strength may be anticipated with high levels of air
entrainment.

SURFACE AESTHETICS

3.2.6 Assessment of Concrete Mixtures / 3.3.1 Open or Closed Units

Apart from stating air-entraining requirements when
necessary, the choice of admixtures should be left entirely to the precaster. However, the architect can request information about admixtures in the concrete
mixtures proposed for the project.

3

Fig. 3.2.8 Different concrete placing conditions.

As a control measure for staining of concrete due to
weathering, it is recommended that maximum water
absorption limits be established. This subject is covered
in greater detail in Section 3.6.5.
A concrete mixture designed for purely structural reasons, or for acid-etched finishes (light exposure of aggregate), is normally fully (continuously) graded, which
means that it contains all aggregate sizes (below a
given maximum) in amounts that ensure an optimum
density of the mixture. However, where aggregates are
to be more deeply exposed by removing the cement/
sand matrix from exposed surfaces, coarse aggregate
in the middle size range may not be able to adhere to
the remaining surface. This may leave too much matrix (sand and cement) exposed, or an uneven distribution of remaining coarse aggregate. To remedy this,
exposed-aggregate panels are commonly produced
using a gap-graded mixture, where one or more of
the intermediate sizes of coarse aggregate are left out.
This leads to a concentration of certain aggregate sizes
in excess of standard gradation limits, which are normally waived for architectural concrete face mixtures,
and improves the panel appearance.
While gap-grading is an established and well-proven
practice, it should not be carried to extremes. This may
cause separation of the paste and aggregates, creating uniformity problems, especially where the mixture
is not deposited close to its final location (Fig. 3.2.8).
The amount of fines, cement, and water should be
minimized to ensure that shrinkage remains within acceptable limits and that surface absorption will be low
enough to maintain good weathering qualities. The
durability of the concrete would normally not be affected by any degree of gap-grading as long as proper
concrete cover is maintained over the reinforcement.
The degree of gap-grading should be based on appearance, but also related to production considerations
and the weathering qualities desired for the specific
exposure of the concrete.
In addition to sample approval and assessment of the
concrete mixture for expected performance, the architect should check the following requirements:
1. D
 ocumentation from the precaster that the con-

Concrete deposited directly in its ultimate location

Concrete flowing through mold before reaching its ultimate location

crete mixture is properly designed for appearance,
strength, durability, and weathering (absorption).
Also that it is suitable for the particular panel
configuration and the anticipated production
techniques.
2. Materials, particularly aggregates, are suitable and
available in sufficient quantities.
3. The precaster has facilities and procedures for uniform batching and proper mixing.
4. The precaster has the facilities, experienced personnel, and established quality control and recordkeeping procedures.

3.3 SHAPE, FORM, AND SIZE
3.3.1 Open or Closed Units
The shape of a precast concrete unit can be an important cost consideration. A major factor is whether the
unit’s shape can be characterized as open or closed.
Precast concrete units should be rigid, to allow for easy
handling; closed units afford this rigidity because of
their shape. Spandrel panels are normally as easy to
handle as closed shapes, although they may occasionally have large returns that require special attention.

ARCHITECTURAL PRECAST CONCRETE

|

111

3

SURFACE AESTHETICS

3.3.1 Open or Closed Units / 3.3.2 Drafts

that open shapes should not be used; their basic weaknesses can be overcome by proper unit proportioning
or by the use of stiffeners or strongbacks.
Combinations of closed and open shapes have better
rigidity, but the cantilevered sections should be proportioned to minimize deflection and tolerance problems.
Close tolerances must be maintained during production and curing to properly match units with open
shapes during installation. The architect may help by
choosing joint details that will minimize deviations (see
Section 4.7.8).
Fig. 3.3.1 Open and closed units.

Fig. 3.3.2 Closed unit.

Fig. 3.3.3 Open and closed units.
Joint

Joint

Joint

Open

Open

Closed

The interfacing of windows and precast concrete
panels is fairly simple in the case of closed shapes because connections and joint details are independent
of site conditions and tolerances, and governed only
by tolerances that relate to the manufacturing for the
two products. In the case of open units and spandrel
panels, the interfacing of windows in the façade will
have to allow for slightly larger and more uncertain
site-construction tolerances. Where window openings
occur between such units, glazing can only be accommodated by a window frame, which considers the appropriate tolerances of the opening.
As described in Section 4.5.2, all panel connections
must allow for minor movements of the panels in relation to the supporting structure. In the case of open
units and spandrel panels, it is important that similar
allowance for movements be designed into the panelto-window connections and the joints between the
concrete and window frame. In the case of closed
panels, these movements are accommodated only in
the joints between the concrete units. Also, windows
in closed shapes can be installed and glazed on the
ground (most often in the precast concrete manufacturer’s plant), which may result in overall cost savings
for the façade and reduce construction schedules.

3.3.2 Drafts
A window unit is a typical example of a closed shape;
the same window unit without the sill or jamb portion
is an open shape. Figures 3.3.1, 3.3.2, and 3.3.3 show
examples of both closed and open unit shapes.
Open units are normally more delicate and may require temporary stiffeners or strongbacks for safe
handling, which adds to cost. Also, some open panels
may be difficult to store without the risk of developing excessive bowing or warping. This does not mean

112

|

ARCHITECTURAL PRECAST CONCRETE

The optimum economy in production is attained if the
panel can be separated from the mold without disassembling the mold. This is accomplished by providing a
draft on the sides of all openings and along the edges
of the panel. In establishing the shape of a panel, the
designer should consider the draft required to strip the
precast concrete unit from the mold, as well as the
draft required to facilitate a specific finish. Generally,
the minimum positive draft for ease of stripping the
unit from a mold is 1:8. This draft should be increased

SURFACE AESTHETICS

3.3.2 Drafts / 3.3.3 Reveals and Demarcation Features

3

Fig. 3.3.4 Draft requirements.

Larger positive draft for narrow member

Minimum positive draft for very narrow faces
and repeated patterns such as ribbed surfaces

for narrow sections or delicate units where the suction
between the unit and the mold becomes a major factor in both strength requirements and reinforcement
of the unit. The draft should be increased to 1:6 for
screen units pierced with many openings, for narrow
ribbed panels, for smooth concrete, and for delicate
units (Fig. 3.3.4). Drafts for ribbed panels should be
related to the depth, width, and spacing of the ribs.
The drafts required for finish consideration are a
function of the shape of the panel, the specified stripping strength of the concrete, the mold release agent
selected, the production techniques, and the desire for
long-term durability. The architect is urged to consult
local precasters for specific recommendations. At areas
where negative draft is required, it may be necessary
to incorporate slip blocks (removable plugs) to aid in
stripping the precast concrete panel from the mold
(Fig. 3.3.4). Vertical sides or reverse (negative) drafts
will create entrapped air voids, which, if exposed, may
be objectionable. Minimizing these surface blemishes
will incur extra cost. Without repetition, mold and production costs increase with negative draft because a
slip block would have to be incorporated with the side
rail and removed with each panel during stripping or
the side rail removed in order to strip the panel. When
the side rail must be removed, dimensional tolerance
becomes a daily variable. Before requiring a negative
draft on the top of a parapet panel, consideration
needs to be given to the roofing or flashing details required for the parapet and the finish. In general, the
greater the draft the architect can allow, the more economical and uniform the finish. A compromise may be
required between the finish and the shape of a precast
concrete unit. A precast concrete unit exposed all the
way around but with good draft for the use of a complete envelope mold is shown in Fig. 3.3.5.

Negative draft removable plug (slip block)

3.3.3 Reveals and Demarcation Features
A reveal or demarcation feature is a groove or a step
in a panel face generally used to create a desired architectural effect, such as separating finishes or mix-

Fig. 3.3.5
Woodrow Wilson School of Public and International
Affairs–Princeton University, Princeton, New Jersey;
Architect: Minoru Yamasaki.

ARCHITECTURAL PRECAST CONCRETE

|

113

3

SURFACE AESTHETICS
3.3.3 Reveals and Demarcation Features

tures (Fig. 3.3.6[a]). Another name for it is rustication
or false joint. Reveals can take vertical, horizontal, diagonal, or curved forms, as well as any combination
of these, and there may be several bands of them on

a building. They can be narrow and delicate or deep,
wide, and bold; they can offer a rectangular profile or
take on any sectional shape desired, such as concave
or triangular.

Fig. 3.3.6 Reveals and demarcation feaures.

Sandblast
Finish
Mix A

Exposed
Aggregate
Finish

Mix B

Sandblast
Finish

Mix A
(b) Scale Reducing

(a) Separate Finishes

Repeated

Real Joint

Real Joint
Single Basic Shape
(d) Double

(c) Single & Repeated

(e) Layered

Typical Flat Mold

Little or no
additional forming costs

(f) Pattern

Typical Reveal Pattern
Back Form
Additional mold and
back-forming costs

Deep Set Reveal
(g) Various Shapes

114

|

(h) Deep Set

ARCHITECTURAL PRECAST CONCRETE

(i) Molds

SURFACE AESTHETICS

3.3.3 Reveals and Demarcation Features

Reveals or demarcation features can add visual interest to a building clad with architectural precast concrete panels while eliminating some of the aesthetic
concerns that develop when planning panel configurations. Used effectively to create shadow lines, reveals
offer the simplest way to reduce the scale of large concrete panels or to keep the visual appearance from focusing on any differences that may occur in texture or
coloration between panels (Fig. 3.3.6[b]).
Reveals can be single (Fig. 3.3.6[c]), double (Fig.
3.3.6[d]), layered (Fig. 3.3.6[e]), or repeated (Fig.
3.3.6[c]). They also can run in patterns (Fig. 3.3.6[f]) or
feature various shapes (Fig. 3.3.6[g]). Deep-set reveals

3

are incorporated in façades to give visual relief and may
require thickened sections (Fig. 3.3.6[h]). Reveals typically measure 1/2 to 3/4 in. (13 to 19 mm) deep and 3/4
to 4 in. (19 to 100 mm) wide, with 45º to 60º beveled
surfaces allowing for ease of stripping, usually 1/16 in.
(1.6 mm) taper per 1/4 in. (6.3 mm). Designers can increase the draft to articulate and manipulate the way
the reveal or panel joint is perceived.
Single horizontal and vertical rustications (Fig. 3.3.7)
tie the precast concrete to similar jointing in the curtainwall system. They also help to reduce the mass of
the large radius precast concrete fascia. Single and
double vertical and horizontal reveals were combined

Fig. 3.3.7
Ahold Information Services
Greenville, South Carolina;
Architect: Smallwood, Reynolds, Stewart, Stewart & Associates Inc.;
Photo: Kieran Reynolds Photography.

ARCHITECTURAL PRECAST CONCRETE

|

115

3

SURFACE AESTHETICS
3.3.3 Reveals and Demarcation Features

Fig. 3.3.8(a)
Adtran Corporate Headquarters, Phase IV
Huntsville, Alabama;
Architect: Cooper Carry, Inc.;
Photos: (a) Gabriel Benzur, Inc.
and (b) Steve Brock.

(b)

Fig. 3.3.9(a) & (b)
Cinedome Theater – The Children’s Museum of Indianapolis, Indianapolis, Indiana;
Architect: Browning Day Mullins Dierdorf.

116

|

ARCHITECTURAL PRECAST CONCRETE

(b)

SURFACE AESTHETICS

3.3.3 Reveals and Demarcation Features

on the structure in Fig. 3.3.8(a). Figure
3.3.8(b) shows a close-up of the crisp reveal details.

3

(b)

A diagonal pattern of 3/4 in. (19 mm) and
3 in. (75 mm) reveals were used to accent
the tipped conical surface (Fig. 3.3.9[a]).
In the 3 in. (75 mm) reveals, a coated aluminum accent strip was inserted at the
intersection of the accent pattern (Fig.
3.3.9[b]). The panels in Fig. 3.3.10(a) and
(b) were match-cast to ensure exact replication of the crossing “wave” pattern cast
in the panels, which extends across the
façade. About 500 pieces, encompassing
90,000 ft2 (8400 m2) of precast concrete,
were used on the project.

Fig. 3.3.10(a) & (b)
Nordstrom Michigan Avenue
Chicago, Illinois;
Architect: Callison Architecture;
Photo: (a) Chris Eden/Callison Architecture.
(a)

ARCHITECTURAL PRECAST CONCRETE

|

117

3

SURFACE AESTHETICS
3.3.3 Reveals and Demarcation Features

The courtyard space in Fig. 3.3.11 indicates the interplay of both the boldly detailed vertical and horizontal
reveals with the building forms. Care had to be taken
when applying drafts for stripping, because many of
the architectural precast concrete loadbearing window-box wall panels have deep returns, and many of
the reveals had to line up around the entire perimeter
of the project. The façade of the 21-story office building in Fig. 3.3.12 consists of two precast concrete panel
types: (1) those at the lower floors with deep negative
grooves, and (2) those at the upper floors with three
horizontal ribs, triangular in section, with infill areas
where a striated pattern is used on a 45º angle.
It is important to remember that a reveal, regardless of
its depth, reduces the structural thickness of the panel.
As a result, when a deep reveal is required, its location
and effect on the panel’s structural performance must
be considered.
A longitudinal reveal has less impact on structural
design than a transverse reveal because the latter decreases the primary bending strength of the panel. A

Fig. 3.3.11
Aurora Municipal Justice
Center, Aurora, Colorado;
Architect: Skidmore,
Owings & Merrill.

horizontal reveal decreases the panel thickness across
the entire width of the panel, while a vertical reveal only
decreases a very small portion of the panel width.
A 3/4 in. (19 mm) reveal has minor consequences on
panel design, and can often be accommodated by
small increases in reinforcement rather than arbitrarily
going to a thicker panel. Deeper reveals of 1 in. (25
mm) or more will generally require a thicker panel, resulting in extra cost.
A good rule of thumb is to have a minimum of 5 in.
(125 mm) of concrete behind a reveal for large panels.
However, it is important to work with both the structural engineer and the precaster to determine how
deep a reveal can be before the panel thickness needs
to be increased. In some cases, reveals with slightly less
than the planned depth of the face mixture can be extremely economical.
Using horizontal reveals within a precast concrete
wall emphasizes floor lines, ceiling lines, or roof lines.
Vertical reveals can express the planning module on
the building’s exterior or its structural rhythm. Diagonal
reveals are almost always a part of a pattern of reveals
applied over the entire structure. Reveals can make
openings within a wall more pronounced or less noticeable. Last but certainly not least, a combination of
techniques can reduce or change the building’s apparent visual scale.

Fig. 3.3.12
Cali International Financial Center, Jersey City, New Jersey;
Architect: Herbert Beckhard Frank Richlan & Associates;
Photo: Herbert Beckhard Frank Richlan & Associates.

118

|

ARCHITECTURAL PRECAST CONCRETE

Precast concrete walls, by their very nature, are made
up of panels or component pieces that are assembled
to create the building’s structure or skin. Those pieces obviously have joints between them, and reveals’
most pragmatic uses come in articulating those fun-

SURFACE AESTHETICS

3.3.3 Reveals and Demarcation Features

damental joints. These joints can be either emphasized
or minimized and hidden by the creative addition of
reveals. Real joints and reveals should have the same
profile. Reveals are generally not caulked. Reveals
other pragmatic uses come in providing drips and/or
small horizontal shelves to protect openings and control moisture movement along the exterior surface of
the precast concrete.
Reveals typically are designed where there are changes in the precast concrete’s surface. For example, a
shift in the panel’s finish from smooth to textured can
be emphasized using a reveal at the point where the
surface texture changes. Reveals also work well where
fundamental materials change within a precast concrete panel, such as from an exposed-aggregate finish to a non-exposed-aggregate finish. Reveals allow a
crisp, clean transition between these different textures,
finishes, or colors.
When the surface of a precast concrete element has
two or more different mixtures or finishes, a demarcation (reveal) feature is a necessary part of the design.
A deep demarcation separates the lightly sandblasted
concrete from the exposed-aggregate center section
of the panel (Fig. 3.3.13). The reveal or demarcation
feature is required to keep the retarder from spreading to adjacent areas. The depth of the groove should
be at least 1.5 times the aggregate size and the width
should be in dimensional lumber increments such as 3/4
or 1.5 in. (19 to 38 mm). The groove should generally
be wider than it is deep so the panel can be stripped
without damaging the mold. A single step in thickness
with a reveal is sometimes used to separate surfaces,
colors, and/or finishes (Fig. 3.3.6[a]).
The importance of the separation provided by a demarcation feature depends on the configuration of the
unit on which the finishes are combined. For example,
a groove or offset is necessary when an exposed-aggregate flat surface is located between widely spaced ribs
with a different surface finish, but not necessary when
a similar flat surface lies between closely spaced ribs.
Proper samples should be used to assess the problem.
The importance of the separation also depends on the
specific types of finishes involved. See Section 3.5.7
for a discussion of finish combinations or variations on
the same panel. Different face mixtures should have
relatively similar behaviors with respect to shrinkage,
to avoid cracking at the demarcation feature due to
differential shrinkage.

3

If a demarcation groove occurs near a change of
section, it may create a plane of weakness (potential
crack) and counter any attempt to provide a gradual
transition from one mass to another. It may be necessary to thicken the section to compensate for the
groove or provide a more rounded groove than would
normally be used. Reglets, window grooves, and false
joints (rustications) will similarly reduce the effective
section of the unit. In some cases, these features may
determine the minimum section thickness required for
the unit.
Lastly, reveals can be placed where there are directional changes in the precast concrete surface, such as between a vertical surface and cornice or bullnose detail.
These elements within a wall design can be emphasized
or de-emphasized through the use of reveals.
Reveals can be much more than a joint or line of demarcation between textures or finishes. Designing reveals in varying shapes, sizes, and depths for a precast
concrete wall can transform what initially might be
considered a mundane, solid surface into a rich texture
of shade and shadow, bringing visual interest to the
building’s façade.

Fig. 3.3.13
The Westin Hotel, Copley Place
Boston, Massachusetts;
Architect: The Architects Collaborative Inc.;
Photo: The Architects Collaborative Inc.

ARCHITECTURAL PRECAST CONCRETE

|

119

3

SURFACE AESTHETICS
3.3.4 Sculpturing

3.3.4 S culpturing
Today, buildings are more sculptural in form, with
a trend toward more organic expressions. There is
greater freedom in the design of the façade. Volumes,
surfaces, lines, and difference in planes are becoming increasingly important in providing architectural
interest. Designers are conceiving of form organically,
generating fluid surfaces in place of rigid structures.
Design focuses on space, structure, and proportion.
Architectural precast concrete provides the designer
with virtually complete sculptural freedom and flexibility in shaping concrete into an articulated structure.

Fig. 3.3.15
461 Fifth Avenue, New York, New York (1989);
Architect: Skidmore, Owings & Merrill;
Photo: Wolfgang Hoyt/Esto Photographics.

Taking advantage of precast concrete’s moldability in creating surface architecture can add considerable aesthetic
appeal to a project.
One of the most important properties of concrete is its
moldability. Concrete is really like sculptor’s clay in an architect’s hands. A wide range of shapes is possible. Concrete
shapes are not limited to volumes enclosed within plane
surfaces: they may also be radiused or rounded.

Fig. 3.3.14
55 Park Place
Atlanta, Georgia;
Photos: George Spence.

Fig. 3.3.16
Moscone Convention Center Esplanade Ballroom
San Francisco, California;
Architect: Gensler/DMJM, Associated Architects (joint venture).

120

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.3.4 Sculpturing

Sculptured panels can produce building façades with
distinctive, strongly modeled elevations having flat interior wall surfaces. High and low relief, straight-line
geometric patterns, and practically any free-form shapes
are possible. The light and shadow effect achieved by
sculpturing the exterior surface produces the major visual effect of precast concrete units. Textures and colors are only of secondary importance when a building is
viewed in its entirety or from a distance (Fig. 3.3.14). The
intricacy and depth of articulation of the façade in Fig.
3.3.15 provides a feel similar to the terra cotta buildings
built at the turn of the century.
Precast concrete panels and precast concrete planter
walls are highly articulated with horizontal bullnose
bands, deep reveals, and strongly expressed horizontal joints (Fig. 3.3.16). A large number of horizontal
setbacks was used in an attempt to de-emphasize the
building size, while still establishing a strong architectural
presence in the complex.

(a)
Fig. 3.3.18(a) & (b)
Jefferson-Pilot Corporate Headquarters
Greensboro, North Carolina
Architect: Smallwood, Reynolds,
Stewart, Stewart & Associates
Photos: Gabriel Benzur.

3

Fig. 3.3.17 Molds–Sculpture, form liner, letering.
(Reversed)
Demarcation strips required
to separate profiles, finishes
and colors

Sculpture – additonal cost
varies depending on depth of
relief, type of lines, degree of
difficulty and repetition

(2046) Cost of letters and layout

Form Liner – additional cost
of liner and minimal form
work

Considering the variety of precast concrete’s sculpting options ensures that its full advantages are used in
designing a façade. These options not only add visual
interest and visually reduce the building’s mass but they
also can customize the building to add personality and
personalization.
Complex shapes and configurations of precast concrete
units may not create a cost premium if sufficient repetition of the unit minimizes the mold costs and where the
sculpturing of the shape aids the unit’s structural capacity. See Fig. 3.3.17 for the effects of sculpturing on mold
costs.
A new headquarters building was needed to complement an existing headquarters, a Romanesque revivalstyle building constructed in 1924 that is adjacent to
the proposed site (Fig. 3.3.18[a]). The original building
was clad in terra cotta tile consisting of 40 to 50 unique
shapes, ornate decorations, and embellishments. The
components and design elements of the existing historical building were analyzed and interpreted into a
system of components appropriate for fabrication into
precast concrete panels. The new 20(b)
story building has a traditionally styled
precast concrete façade with elaborate detailing, colored and formed to
resemble the weathered terra cotta of
the original structure. Each piece was
designed with false joints, which gave
the illusion of hand-carved stone and
diminished the effect of color variations. By using precast concrete panels,
a more apparent degree of depth, detail, and richness of
the ornate designs was achieved (Fig. 3.3.18[b]).

ARCHITECTURAL PRECAST CONCRETE

|

121

3

SURFACE AESTHETICS
3.3.4 Sculpturing

Fig. 3.3.19
Citicorp Tampa Campus
Tampa, Florida;
Architect: HKS Architects Inc.;
Photo: Michael Lowry Photography.

els is an added expense, it may be necessary to reduce
the weight of the panels. In certain light finishes, such
as acid-etched or light sandblast, a temporary shadowing of ribs may be visible. In units with ribs in only one
direction, the dimension in the other direction might
either be shortened or strengthened by using ribs on
the back. A panel may have reasonable stiffness in the
vertical direction but be weak horizontally. The precaster may choose to improve the structural strength
by incorporating a concrete rib.
In most cases, dimensions of ribs will be determined as
part of the architectural features of the units. Minimum
dimensions are determined by design; and practical
considerations are treated in detail in Section 4.2.9.
The visual impact made by the relief sculpture depends mainly on two factors: profile and lighting. The
profile or cross-section should consist of strong elements with edges that produce well-defined highlights
and shadows. Surfaces that flow smoothly into each

Sculpturing may increase structural strength of the
precast concrete units and, therefore, simplify handling. The panels should be shaped for sufficient stiffness in the direction of handling-induced stresses.
Precast concrete panels molded around windows are
often set forward of the glazing, adding stiffness and
giving sculptural form. Sculpturing also may increase
the depth-to-span ratio through ribs or projections in
either direction of a unit (Fig. 3.3.19 and 3.3.20). The
depth of panels is defined as the thickness and the
span would be either the height or width of a unit.
With sufficient panel repetition, and where the depth
and volume of the projections do not exceed the optimum required for handling, there should be no cost
premium beyond the cost of the added volume of
materials.
The projections do not have to be continuous or
straight, but may be overlapping or curved. Projection
design should avoid creating a weak section within the
units. Projections may not add to structural capacity
when they are interspersed between weaker sections.
Ribs may be part of the architectural expression or,
where flat exterior surfaces are required, ribs may be
added to the back of panels for additional stiffness.
Although backforming for the rib on the back of pan-

122

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 3.3.20
100 South Fifth Street
Minneapolis, Minnesota;
Architect: OPUS Corporation;
and Hellmuth, Obata &
Kassabaum, P.C.

SURFACE AESTHETICS

3.3.4 Sculpturing

3

(a)

(b)

(c)

(d)

(e)

(f)

(g)

Fig. 3.3.21(a - g) Various patterns integrated into formwork.
Nimitz-MacArthur Pacific Command Center, Oahu, Hawaii; Architect: Wimberly Allison Tong & Goo Design; Photos: Gary Hofheimer Photography.

other should be avoided. A bold treatment is most effective with subtle or gradual changes in the profile.
These limitations are very important for the cross-section, but do not apply to the front or elevation view
of the design. If it is possible to control lighting, make
sure it plays across the relief from the side rather than
straight from the front.
Relief sculpture can be enhanced by contrasting
surfaces on projecting elements with textures on the
background.
The viewing distance of the surface should be considered when deciding on the scale of the relief. As a
rough guide, design elements smaller than about 1/300
of the viewing distance are difficult to “read” and tend
to get visually lost in their surroundings.
The use of precast concrete in public art applications
is growing in popularity. A wall with creative images

reduces the visual scale of the panels and turns the
wall into a work of art.
An example of relief sculpture is shown in the panels for the building in Fig. 3.3.21. The design intent
of the two-toned, colored precast concrete façade
was to capture the Hawaiian architectural style as it
relates to the natural elements, such as earth, water,
mountains, and sky. These elements are picked up in
the different levels of the façade in the form of stone
textures, waves, multi-level reveal patterns, three dimensional pineapple leaf patterns, fluted mullions,
heavy cornices, dentils, bullnoses, ribs, and the navy
globe symbol that was sculptured and integrated into
the formwork (Fig. 3.3.21[a] to [g]). All of these were
accomplished using various types of forming materials
best suited for a particular situation, including steel,
fiberglassed wood, urethane liners, milled plastic, and
sprayed fiberglass.

ARCHITECTURAL PRECAST CONCRETE

|

123

3

SURFACE AESTHETICS
3.3.4 Sculpturing

Fig. 3.3.23
Tropicana Casino Addition
Atlantic City, New Jersey;
Architect: Ellerbe Becket formerly Welton Becket & Associates;
Photo: Ellerbe Becket.

Figurines for the murals were cast in five separate
molds 4 ft × 7 ft 6 in. (1.2 × 2.3 m) and set into an 8 ft
× 15 ft (2.4 × 4.5 m) wood mold before casting the final
panel (Fig. 3.3.23). Figurine panels are 6 in. (150 mm)
thick and the figures project up to 11/2 in. (38 mm) from
the flat areas.
Fig. 3.3.22
Level 3 Communications
Needham, Massachusetts;
Architect: HLW-Thomson Design;
Photo: Peter Paige.

Between thin brick–clad precast concrete
panels reside feature panels of architectural
precast concrete with the stylized design of
a computer circuit board (Fig. 3.3.22). It is replete with stainless steel “chip” mounted at
the center of the panel and acts as an indirect
lighting fixture, which, at night, splays light
across the molded surface of the circuit board
panel, illuminating each raised feature. From
the “chip” emanates circuit details that flow
across the surface of the panel to align with
either the 4 ft (1.2 m) brick panel module
horizontally or vertically to stylized “transistors” cast as entablatures above and below
the chip.
Fig. 3.3.24
The Parking Gallery, Reno, Nevada;
Vicki Scuri, Sculptor;
Photo: Vicki Scuri.

124

|

ARCHITECTURAL PRECAST CONCRETE

The butterfly sign at the entrance to the parking
structure was replicated in the precast concrete spandrel panels (Fig. 3.3.24).

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

3.3.5 B
 ullnoses, Arrises and Radiused
Precast Concrete
The bullnose offers a useful tool with which architects
can increase visual interest by adding dimensionality
and allowing the design to avoid simple flat surfaces.
Three-dimensional pieces that extend from a flat surface change the reading and proportion of that surface. The light and shadow variations achieved with a
bullnose produce a major visual impact and contrast
when a building is viewed from a distance. Also, shadows cast by a horizontal bullnose profile create strong
lines that reduce the apparent height of the structure.

3

Fig. 3.3.25 Bullnoses and Arrises.

(a) Typical bullnose

(b) Double bullnose

(c) Elliptical bullnose

(d) Bullnose with
reveals at top
and bottom

(e) Partial bullnose

(f) Bullnose featuring return

Bullnoses may be designed in a variety of sizes. As the
bullnose increases in size, it adds weight and cost to the
panel, primarily due to the expense of the mold. Here
are some key points to remember when designing bullnose components. For each item, the letter corresponds
to Fig. 3.3.25 that shows the discussed aspect:
(a) T he basic bullnose is 180 degrees, or a half-circle.
(b) M
 ultiple bullnoses can be used within a panel.
(c) The bullnose can be elliptical.
(d) A
 reveal (rustication) may be placed at the intersection of the bullnose and the panel field to accentuate the bullnose. The reveal may also be used
to separate dissimilar mixtures and/or finishes.
(e) A partial bullnose may be designed.
(f) A return may be incorporated with the bullnose.

(g) Cove bullnose

(h) Partial convex bullnose

(g) A half-circle cove.
(h) The convex bullnose may be partial.
(i) The bullnose may feature a finish transition (similar
to [d]).
(j) Arrises (shapes) may be rectilinear or pointed. They may
protrude or be inverted similar to items (a) through (i).
They also may be combined with bullnoses.

(i) Bullnose featuring a
finish transition

The architectural features on the panels in Fig. 3.3.26
include bullnoses with three different radii (61/2, 3, and
11/2 in.) [163, 75, and 38 mm]. Also included are 3/4 in.
(19 mm) deep horizontal reveals. Light and deep sandblast finishes gave the panels two complementary colors
and textures.
The panels in Fig. 3.3.27 use two finishes to create a
banded appearance that accentuates the horizontal expression. A heavy bullnose is incorporated into the spandrel panels to create a window sill and reinforce the horizontality of the wall, as well as add texture to the wall.

(j) Arrises can take many shapes

ARCHITECTURAL PRECAST CONCRETE

|

125

3

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

Fig. 3.3.27
Parkway Plaza, Greenville, South Carolina;
Architect: Urban Design Group Inc.;
Photo:Urban Design Group Inc.

Fig. 3.3.26
Pioneer Place Block 50, Portland, Oregon;
Architect: E.L.S. Architecture and Urban Design.

The large and small diameter bullnoses over the windows in Fig. 3.3.28 do not extend across the entire panels.
The profile of the upper-floor white concrete spandrels in
Fig. 3.3.29, however, includes a large diameter bullnose
running the length of the floor line. The bullnose provides
a graduation between light and shadow in the powerful
desert sunshine, where shadow-lines are extreme: a deep
darkness contrasted with bright sunshine.
Adding bullnoses along each window ledge added vari-

Fig. 3.3.28
Westwood Executive Center
Westwood, Massachusetts; Architect: Sasaki Associates Inc.;
Photo: Desroches Photography.

126

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

3

Fig. 3.3.29 City Center West, Las Vegas, Nevada; Architect: Urban Design Group Inc.; Photo: Urban Design Group Inc.

(a)

ety to the design, improved the proportions of the façade,
and created changing light patterns during the day (Fig.
3.3.30[a]). As part of the same project, large bullnose components added to the parking structure provided a distinctive visual element that tied it to the other buildings in the
project without mimicking their exact look (Fig. 3.3.30[b]).
The fundamental appeal of the bullnose form in precast
concrete design comes from its ability to visually reproportion an uninteresting, flat surface. The bullnose can
also be used to develop more complex forms in combination with bullnose shapes of different radii or in combination with convex, concave, or flat sectional shapes.

Fig. 3.3.30(a) & (b)
Hyperion Wastewater Treatment Facility, Los Angeles, California;
Architect: Anthony J. Lumsden & Associates; Photos: Anthony
J. Lumsden, former, corporate director of design for DMJM.

(b)

Radiused shapes are generally more costly than flat
surfaces, because of the additional work required to
manufacture the mold and to place the reinforcement,
connection hardware, and concrete. See Fig. 3.3.31
for a discussion of factors affecting production costs
for radiused units and Fig. 3.3.32 for applications of
various radiused elements.
Some architectural forms that are not flat can be difficult to achieve. However, forms that are cylindrical in nature, forms that have surfaces generated from a sectional
shape that are consistent throughout the length of the
mold, are simple to form. These forms that are consistent
sectionally allow multiple castings and are economical to
fabricate because attached pieces are identical and easy
to install. Molded shapes that have curvatures about both
axes are difficult to fabricate, difficult to install, and have
limited repetitive applications.

ARCHITECTURAL PRECAST CONCRETE

|

127

3

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

Fig. 3.3.31 Molds.

Fig. 3.3.32 Examples of radiused units.

Soften a building return; partial Mold G
A: Typical mold

B: Bullnose — some additional form costs

C: Inverted arrise — additional formwork
plus labor to daily remove and replace
back pans

Form a bullnose; Mold B

Photo: Gabriel Benzur.

D. Column cover — more complex
formwork plus additional labor daily to
remove and replace the back pan.

Enclose a structural
member; Mold D

E: & F: Gradual radius — additional forming
and additional labor to back finish

128

|

ARCHITECTURAL PRECAST CONCRETE

Create a building
extension; Mold
E&G

Photo: Gabriel Benzur.

G: & H: Extreme radius — more complex
formwork with sequential back pans that
must be removed and replaced daily.
Also, casting time takes longer and some
back finishing is required

Form a circular window or opening

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

3

Create a combination
such as an arched roof
member including a
beam surround
Form a concave building corner;
Mold F & H

Create an undulating surface using a combination of
Molds E & G; Molds F & H

Photo: Rick Alexander & Associates, Inc.

Form convex building
corners; Molds E & G

Form arches

Photo: ©Anton Grassel.

Form an inverted arrise;
Mold C

Create circular patterns with reveals

ARCHITECTURAL PRECAST CONCRETE

|

129

3

SURFACE AESTHETICS

3.3.5 Bullnoses, Arrises and Radiused Precast Concrete

Fig. 3.3.33 Examples of cornices and eyebrows.

Finish A

Finish B
Typical flat mold

Finish A

Additional
forming cost
(a) Contemporary
Typical cornice
Intermittent
support

Continuous
void form

Complex mold
plus back pan
to be removed
and replaced
daily
(b) Traditional
Ornate cornice

Reveal

Rear

(c) Doubled

(d) with Reveal

(e) with Coping

rear view

Haunch
(2 per panel)
(f) Special haunch on panel

(g) Special Structural Supports

Dentils

Front view

(h) with Dentils
(i) Eyebrows

130

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.3.6 Cornices and Eyebrows/3.3.7 Edges, Corners, and Returns

3.3.6 C ornices and Eyebrows
A cornice, as an element of the façade, has three primary functions:

different purposes. Options in Fig. 3.3.33 include cornices that are:
• Made to look contemporary (a).

• It provides “the termination” of the vertical spread
of the building. It’s the top, pure and simple.

• Made to look traditional (b).

• It provides a balance and proportion to the entire
façade, acting as a counterweight to the aesthetically heavier base of the building.

• Incorporated with a reveal (d).

• When cantilevered 12 to 24 in. (300 to 600 mm)
away from the plane of the main façade, the cornice serves a function, acting as a rain shield for the
upper floors and helping to minimize dirt streaking
and water stains.
A cornice consists of a horizontally projecting overhang comprising multiple surfaces, planes, and profiles
with thousands of variations. It usually is located along
a parapet or at the top of a given plane. The cornice
crowns or finishes the part to which it is affixed.
When used as a horizontal projecting element that is
not situated at the building’s top, the traditional cornice-shaped element becomes an “eyebrow” or “shelf”
offering additional aesthetic proportion and definition
to the entire façade. If one believes that buildings encompass the three basic parts of base, middle, and top,
these eyebrows can define the transition from one part
to another or provide the transition from one type of
building element to another, such as with articulated
column capitals.
Today’s architectural vocabulary also might employ
this device as a light shelf (reflecting light) or shadow
maker, as it can shade windows from sun and rain, reducing energy costs and other internal shading needs.
Either will develop interesting and ever-changing light
and dark patterns on the surfaces below.
Whether used as the top piece or as an eyebrow, the
architectural precast concrete cornice shape offers architects a multitude of design possibilities. For instance,
a cornice easily can cantilever past the structural roof
slab or project away from the façade’s plane without
needing complex additional support. When the design
emphasis articulates a “heavy, large expression at the
top of the building,” precast concrete pieces can accomplish both the aesthetic and the functional needs
of this concept.
Cornices can be used in a variety of styles and combined with several different components to achieve

3

• Doubled to create even more design interest (c).

• Cast with a void form to reduce weight or with coping on the crown ([e] and [g]).
• Created so large that it requires special support.
For instance, a steel structure may require bracing
to prevent rotation of structural members ([f] and
[g]).
• Incorporated with dentils (h).
Mold costs for either cornices or eyebrows will depend on the degree of complexity and the size of the
projection (i). Both cornices and eyebrows (i) may be
continuous, interrupted, arched, or peaked. Dissimilar
finishes may be used on adjacent surfaces (a). When
detailing cornices attention needs to be paid to the
window washing system to avoid damage to the cornice. A montage of various cornice styles is shown in
Fig. 3.3.34.

3.3.7 Edges, Corners, and Returns
Each individual project requires special attention to
the design and detailing of its corners to create optimum appearance, jointing, and economy. For this reason, corner detailing should be decided early. Economy
results when the building elevations are designed from
the corners inward, using typical panels and avoiding
special-sized end or corner pieces. Typical corner treatments, such as a mitered edge or a 12 in. (300 mm)
corner return, usually influence all corner pieces for the
project. Isometric sketches of the building that show
panel layout will help define areas where corner panels
are needed.
All edges of precast concrete units should be designed
with a reasonable radius, chamfer, or quirk, rather
than leaving them as sharp corners. This is particularly
important where the panels are close to pedestrian or
vehicular traffic. The size of the edge’s radius should
be discussed with the local precaster. Determining the
optimum size depends on the selected aggregate size,
mold materials, and production techniques. When the
edge is sharp, only fine aggregate collects in these

ARCHITECTURAL PRECAST CONCRETE

|

131

3

SURFACE AESTHETICS
3.3.6 Cornice and Eyebrow

Fig. 3.3.34(a – h) Montage of various cornice styles.

(a)

(b)

(d)

(e)
(c)

(g)
(f)

(h)
Photos: (a) Gabriel Benzur Photography; (b) Cathers & Associates, Inc.; (c & d) Brian Gassel/TVS; (e) Grodon Dilgore; (g) James Oesch
Photography/Donnally Vujcic Associates, L.L.C.; (h) Lambros Photography Inc.

132

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.3.7 Edges, Corners, and Returns

Fig. 3.3.34(a – h) Photo credits

3

Fig. 3.3.36 Quirk miter dimensions (in inches).

Fig. 3.3.34(a) Douglas County Courthouse, Douglasville,
Georgia; Architect: Cooper Carry & Associates.
C — joint

Fig. 3.3.34(b) 2201 Renaissance Corporate Center,
King of Prussia, Pennsylvania; Architect: Cathers & Associates
Inc.
B

Fig. 3.3.34(c) & (d) 2300 Lakeview Parkway Building,
Alpharetta, Georgia; Architect: Thompson, Ventulett and
Stainback Associates (TVS).
Fig. 3.3.34(e)
Resurgens Plaza, Atloanta, Georgia; Architect: Smallwood,
Reynolds, Stewart, Stewart & Associates, Inc.

A
quirk

B
Quirk miter dimensions
90º

Fig. 3.3.34(f) St. Charles County Courts Administrative
Building, St. Charles, Missouri; Architect: Sverdrup Facilities
Inc.

A
quirk

Fig. 3.3.34(g) Presidents Park, Hendon, Virginia; Architect:
Donnally, Lederes, Vujcic, L.L.C.
Fig. 3.3.34(h) Leaf North America Corporate Headquarters
Lake Forest, Illinois; Architect: Loebl Schlossman &
Hackl/Hague Richards.

locations and this weakens the edge. Voids also occur due to the interference of larger aggregate. Sharp
corners chip easily, both during handling and during
service on the finished building. Chamfered or radius
edges also mask minor alignment irregularities of the
precast concrete panels. For typical edge details, see
Fig. 3.3.35.

C — joint

B

Fig. 3.3.35 Typical edge details.

45º

B
Quirk miter dimensions
45º

A
quirk
22-1/2º

A
quirk

Mitered corners without quirks are difficult to manufacture and erect within tolerances that are acceptable from either an appearance or a jointing standpoint. Concrete at mitered corners cannot be cast to a
sharp 45º point because of the size of the aggregates.
Therefore, this edge must have a cut-off or quirk (Fig.
3.3.36). A table showing the recommended size of the
quirk return for different panel-joint sizes is included in
Fig. 3.3.36. The size of the quirk return should never
be less than 3/4 in. (19 mm), nor less than 1.5 times the

ARCHITECTURAL PRECAST CONCRETE

|

133

3

SURFACE AESTHETICS
3.3.7 Edges, Corners, and Returns

Fig. 3.3.38 Typical corner and return details.

Slab
Edge

Miter Corner

Slab
Edge

Corner panel (return)
included on adjacent
panel

Butt Corner

Slab
Edge

maximum size of the aggregate used in the concrete
mixture. Normally a 3/4 to 1.5 in. (19 to 38 mm) quirk
will read as a well-defined edge on the corner of the
building. A well detailed and fabricated miter and a
quirk miter are shown in Fig. 3.3.37.

How the precast concrete is being used and the type
of panel that is turning the corner determines how the
building corners and major component edges will be
designed. Figure 3.3.38 shows typical corner and return details.
Flat panels, used either to visually define the dimensioned mass of building-block elements or to create
flat or curvilinear planar surfaces, are treated differently than panels with heavily articulated horizontal
treatment using deep relief reveals or profiles. Visual
focus at the corners often is part of a design approach
where the wall plane is stopped or interrupted at the
building corner or the corner is emphasized to define
the building form.

134

|

ARCHITECTURAL PRECAST CONCRETE

Note:
Vertical casting may
cause variations in
finish from the face
down side.
Some finishes may
not be consistent over
12in. (300mm) high.

Min. Rad.
Vertical Cast Corner

Joint
Angle as Design

Typically up to
3’0” each side

Even with good-sized quirk returns, a mitered corner
may cause the panels to converge at the top, bottom,
or center, depending on the vertical configuration of
the panels. If the building design demands corners
with mitered edges, the architect is urged to specify a
mockup of the two initial corner panels at the precast
concrete plant before approving the panels and releasing the balance for production.

12" Max.

Adjacent
Panel

Slab
Edge

Note:
The cast corner is
typically used when
one or both legs of
the corner exceeds
the limitations of a
formed corner.

Min. Rad.

Joint

“V“ Mold Corner

Adjacent
Panel
Slab
Edge

Joint
As Designed

Fig. 3.3.37
Times Publishing Company, St. Petersburg, Florida;
Architect: TRO Jung/Brannen Associates, Inc.;
Photo: George Cott/Chroma Inc.

Note:
The length of a cold
jointed corner is a
matter of handling
and shipping
practicalities.

Cold Joint
(Caulk Optional)
Two-Stage Casting
Contact precaster to determine the best solution and
cost implication of detail.

SURFACE AESTHETICS

3.3.7 Edges, Corners, and Returns

Quirk miter corners, channel shapes with returns, and
two-stage precasting all have achieved the desired result,
often with modifications contributed by the precaster
that benefit both the finished design and the budget.
The projects in Fig. 3.3.39(a) through (j) illustrate various corner treatments. Providing a visible expression of
the building panel or unit width as it turns the corner
gives substance and thickness definition to the material. The heavily rusticated panels in Fig. 3.3.39(a) are
thickened flat panels with a deep quirk return, in the
range of 3 in. (75 mm), to economically create a perceived dimension in the panel as it turns the corner.
A similar deep-quirk miter is used in Fig. 3.3.39(h) for
both a return dimension and emphasis on the corner.

3

When the panel thickness itself is to be expressed, a
butt joint around the corner at the back of the panel is
sufficient without a return. In Fig. 3.3.39(b), one-piece,
channel-shaped covers with relatively shallow returns
were used to imply a thickness somewhat greater than
the actual panel thickness. In this case, well-executed
“sharp” corners in the casting add to the solid appearance. In Fig. 3.3.39(c), two-stage cast covers with deep
returns were used for solid-block treatments of the
piers. The two-stage precasting allows a unique castin-surface texture to be carried around the corners of
the blocks to a much deeper dimension. Minimal corner dry joints without quirks complete the appearance
of a solid unit.

Fig. 3.3.39(a – j) Various corner treatments by Thompson, Ventulett, Stainback & Associates (TVS), Architects; Photos: Brian Gassel/TVS.
(b)

(a)

(d)

McCormick Place Expansion,
Chicago, Illinois
1335 Windward Concourse,
Atlanta, Georgia

Merrill Lynch Campus,
Denver, Colorado

55 Farmington,
Hartford, Connecticut

(c)

When the wall surface is treated as a planar surface,
smooth or articulated, the goal usually is to make the
transition around the corner as smooth and continuous as possible. This often involves moving the joint
from the prominence of a corner location, especially
with smooth panel faces. Figures 3.3.39(d), (e), and
(f) illustrate situations where a solid-appearing corner
is preferred. This allows the smooth surface to carry
around the corner uninterrupted and without calling
attention to the corner.
The flat wall plane in Fig. 3.3.39(d) changes direction
without corner articulation by using a shallow angle return for single casting and placing the joints at the back

ARCHITECTURAL PRECAST CONCRETE

|

135

3

SURFACE AESTHETICS
3.3.7 Edges, Corners, and Returns

(e)

(f)

McCormick Place
Expansion,
Chicago, Illinois

angles where they are de-emphasized. The shallow
corner angle allows a solid cast edge. In Fig. 3.3.39(e),
the returns on the column enclosures are not intended
to imply a material thickness of panels or blocks but to
provide a planar wrap of the surface forming the corner. The joint is located at the pier’s center to provide
a linear continuation of the element above the pier. A
two-stage casting and dry-joint corner in Fig. 3.3.39(f)
provide an abrupt but continuous return of the plane
of the curvilinear sweep to the building face.
On the other hand, horizontal spandrel panels with
deep rustication, or a contained sculptured profile,
such as that shown in Fig. 3.3.39(g), require a miter to
carry that panel profile accurately around the corner.
The minimal quirk-miter corner in the spandrel panels
provides an uninterrupted planar return of the profiled
wall-panel surface without corner articulation. A twostage casting at the cornice provides a similar planar
wrap around the top. These sharp corners at the top
emphasize the building corner’s large-scale articulation. Details closer to pedestrians use the quirk corner
for more finished detail.
Special corner treatment often is used to emphasize
the corner to define the intersection of the wall planes
with detail as shown in Fig. 3.3.39(h) and (i). The degree
of detail complexity and richness, which relates to budget, determines the casting technique. The depth and
shadowing of the façade on the parking structure, Fig.
Metropolitan Life Building, Atlanta, Georgia

136

|

ARCHITECTURAL PRECAST CONCRETE

ADP Building, Atlanta, Georgia

(g)

SURFACE AESTHETICS

3.3.7 Edges, Corners, and Returns

3

(i)

Bell South Campanille, Atlanta, Georgia
Glenridge Highlands One,
Atlanta, Georgia

(j)

(h)

Carrillon Building,
Charlotte, North Carolina

3.3.39(h) is continued in the detailing of the pier corners
by using deep quirk miters to accent the corners with
the shadow created by the larger quirk. Separate corner
pieces Fig. 3.3.39(i) are used to interrupt the wall plane
at the corner, emphasizing the corners by vertically articulating the larger multistory lobby space at the base
from the typical floors above. The expense of using a
special corner mold to construct the elegant corner panels on the granite veneer-faced precast concrete panel
building was justified based on a value engineering
study. It allowed a more intricate corner detail and the
flat areas of granite were able to be panelized.

Figure 3.3.39(j) represents a stronger corner treatment. Solid precast concrete corner elements are used
to define the corners and the transitions from curtain
wall to precast concrete panels at the corners. At the
macro scale, these solid pieces have their corners articulated with quirks to add visual interest and to facilitate corner quality control. Originally planned as
returns and channel spandrel covers, they were cast
solid at the precaster’s suggestion for economy and
detailing simplification.

ARCHITECTURAL PRECAST CONCRETE

|

137

3

SURFACE AESTHETICS

3.3.7 Edges, Corners, and Returns / 3.3.8 Returns in Relation to Finishes

An alternate to mitering is the use of a separate small
and simple corner panel to add interest to the façade.
Special corner pieces can be cast by using modified
standard unit molds, which are part of the master mold
concept discussed in Section 2.3.3. If the size of the
project or the available time constraints warrants multiple molds, a separate corner mold is recommended.
On a high-rise building, the cost of a corner mold and
the handling of an extra piece may offset the modification costs of the master mold and/or be justified by the
additional flexibility in erection tolerances. Separate
corners also may be advantageous in providing similar
orientation of corner surfaces for matching finishes or
the corner pieces may economically be designed and
produced as part of one of the adjacent typical panels. Matching finishes can be very difficult when one is
on a down-face and the other is on a return (vertical)
face.
The variations in the overall length of a building elevation, assuming that these stay within stated tolerances, may either be accommodated in the joints or in
the design of the corner pieces (see Section 4.6). Due
to the reduced size of the corner panels, they will normally undergo less thermal movement and can therefore tolerate greater joint width variation.

(b)

(c)

Figure 3.3.40 displays some of the various shapes of
column covers that can be made with butt or mitered
joints, including (a) two L-shaped units, (b) two Lshaped units with miters, and (c) two U-shaped units.
There are, of course, many other combinations that
can be used to accommodate isolated columns, corner
columns, and columns in walls.
Sequential and monolithic corner (return) molds are
more costly than mitered molds (Fig. 3.3.41). A separate corner panel often requires an additional mold.
The treatment of building corners, as well as smaller-

138

|

ARCHITECTURAL PRECAST CONCRETE

Typical flat mold
Very little form expense

Single cast

Mitered corner
Sequential

Mitered corner

1st cast

2nd cast

Both add considerable form cost

scale, building-component corners, is critical to the
final perception of the architecture. The corners are focal points where wall planes and materials change or
continue. The corners outline and define the form of
the building and the corners are where the light falling
on the surfaces of the façades transition dramatically.

3.3.8 Returns in Relation to Finishes
The precast concrete unit’s finish should be considered before its shape is finalized. Many finishes cannot
be achieved with equal visual quality on all faces of
the unit.

Fig. 3.3.40 Column covers with returns.

(a)

Fig. 3.3.41 Corner – return molds.

The reasons encompass factors such as mixture proportions, variable depths (and pressures) of concrete,
and small differences in consolidation techniques, particularly in the case of intricate shapes with complex
flow of concrete. The affect of gravity during consolidation forces the large aggregates to the bottom and the
smaller aggregates, sand, and cement content upward.
Consequently, the down-face in the mold nearly always
will be more uniform and denser than the returns or
upper radius of curved panels. This usually should be
dealt with during the sample-approval process.
Consolidation of the concrete results in a more or less
uniform orientation of the aggregate, with the flat,
long portion horizontal to the bottom of the mold.
On returns and the upper radius of curved panels, the
sharp angular points of the aggregate will show upon
exposure. This can give returns greater than 12 in. (300
mm) a finished texture distinctly different from that of
the down-face (Fig. 3.3.42).

SURFACE AESTHETICS

3.3.8 Returns in Relation to Finishes / 3.3.9 Two-Stage or Sequential Precasting

Fig. 3.3.42 Exposure variances caused by aggregate.

Different exposures

With deep returns, a more uniform finish is obtained
with an exposed-aggregate finish. When an exposedaggregate finish is specified, concrete mixtures with
aggregates that are reasonably spherical or cubical
should be chosen to minimize differences between
down-faces and returns. For panels with large returns,
or other situations where variations in appearance must
be minimized, the two-stage or sequential production technique should be used if feasible (see Section
3.3.9). Otherwise, concrete mixtures should contain a
continuous-graded coarse aggregate and an ASTM C
33 sand. Exposure of aggregates should be medium to
deep with minimal color differences between mixture
ingredients.
If the units are cast so that surfaces with identical
orientation on the building are cast in similar positions,
small differences in finishes between areas cast facedown and as a return should be acceptable. When this
mold and casting approach is taken, the choice of finish in relation to the configuration of the unit becomes
one of determining the acceptable differences in textures between down-face and varying return surfaces.
This is best judged by observing sample panels from
distances and positions simulating the viewing positions for the finished building.
Sculptured panels, channel panels, and panels with
deep returns may have visible air voids on the returns.
These air voids, or “bug/blow holes,” become accentuated when the surface is smooth, acid-etched, or lightly
sandblasted. If the air holes are of a reasonable size, 1/8
to 1/4 in. (3.2 to 6.3 mm), it is recommended that they
be accepted as part of the texture. Filling and sack-rubbing could be used to eliminate the voids. However, this
procedure is expensive and may cause color differences.
Samples or the mockup panel should be used to establish acceptable air void frequency, size, and distribution.
The architect should accept a small difference between
molded and non-molded surfaces, but the orientation
of these surfaces should be determined on the draw-

3

ings or specifications and be consistent throughout the
project. The architect should also limit the choice of
aggregates and finishes to those that lend themselves
to a reasonable matching of the molded surfaces with
a hand-finished open surface. Therefore, a smooth finish or one with a light exposure is not appropriate. The
open surface may have to be seeded with the larger
aggregates as part of the finishing process. When the
shape of the units precludes any logical demarcation of
formed and uniformed surfaces, closed molds cast in a
vertical position may be the only answer. This will normally require complicated and expensive molds and/or
deep castings.

3.3.9 Two-Stage or Sequential
Precasting
Any portion of a panel cast in a vertical position will
not show the same concentration or positioning of aggregate as the flat surface. Panels with large or steep
returns (such as channel column covers and some
spandrels) may be cast in separate pieces in order to
achieve matching high-quality finishes on all exposed
faces and then joined with dry joints (Fig. 3.3.43[a]
and [b] and 3.3.44[a] and [b]). This method of casting enables all faces to be cast face-down with the
same aggregate orientation and concrete density using
Fig. 3.3.43(a) Separate casting stages of large returns.
Projecting reinforcement

Second cast in mold “A”

Cradle to
support
returns

Third cast with projecting reinforcement cast in,
lapped with reinforcement in face, not shown

Cold joint
Final Profile

ARCHITECTURAL PRECAST CONCRETE

|

139

3

SURFACE AESTHETICS

3.3.9 Two-Stage or Sequential Precasting / 3.3.10 Overall Panel Size

Fig. 3.3.43(b)

Fig. 3.3.45
Inter-Industry Conference on Auto Collision Repair (I-CAR)
World Headquarters
Hoffman Estates, Illinois; Architect: Loebl Schlossman
& Hackl; Photo: David Clifton/Loebl Schlossman & Hackl.

conventional precast concrete forming methods; backforming is not required. Also, a combination of face
mixture and backup mixture can be used, rather than
a 100% face mixture.
If this is the indicated production method, attention
should be paid to suitable corner details and reinforcement at the dry joints. Reinforcement is left sticking
Fig. 3.3.44(a) Alternative casting approaches.
Welded
wire
fabric

(a)
(b)

Rotate
panel-place
in second form

Backup mix

Face mix

Face mix

Quirk miter
corner joint
First cast

(c)

Second cast

3/4"

Drip

Fig. 3.3.44(b)

140

|

ARCHITECTURAL PRECAST CONCRETE

out of the return piece to ensure that the two pieces
are adequately tied together. Although the dry joint
may not show with certain mixtures and textures, precasters recommend a groove or quirk to help mask the
joint. Sometimes precautions may be necessary to ensure watertightness of the dry joints. The main disadvantage of two-stage precasting is that two or three
separate concrete placements are necessary to complete a panel. Figure 3.3.45 shows panels that were
sequentially cast because of large returns.

3.3.10 Overall Panel Size
Panel geometry, referred to in a general sense as shape
details, that does not affect the architectural concept,
can be a major influence on both fabrication economy
and engineering requirements. The significant shape
details are overall size and configuration. The visual
characteristics of a panel are determined by the architect. The size of the individual elements and the exact
details of the panel geometry should be determined in
consultation with the structural engineer and the precaster. Thus, both the architect and the structural engineer should be familiar with good production practice
as well as production and erection capabilities of the
probable fabricators. For overall economy, early coordination of design and erection are essential.
Because many of the manufacturing, handling, and
erection costs are independent of the size of a piece,

SURFACE AESTHETICS

3.3.10 Overall Panel Size

making panels larger can substantially reduce the total
cost of the project. The larger the panel, the smaller
the number of pieces required for enclosure, which
means less handling and lower erection costs.
Hoisting a precast concrete unit constitutes a significant
portion of the cost of precast concrete. The cost difference in handling a large unit, rather than a small one,
is insignificant compared to the increased square footage of the large unit. In addition to providing savings on
erection costs, large panels provide secondary benefits
of reduced amounts of caulking (fewer joints), better dimensional controls, and fewer connections. Thus, large
units are preferable unless they lack adequate repetition, require high forming costs, or incur cost premiums
for transportation and erection. In high seismic areas,
very large panels may not be desirable because larger
dimensional deformations in the supporting structure
must be accommodated. In these areas, normal practice
is to use panels that are one story in height and seldom
span more than one structural bay. Where desired, the
scale of large panels may be reduced using false joints
(rustications).
The wide variety of sizes, shapes, materials, and functions that can be incorporated into a precast concrete
unit makes it difficult to present a size chart. Section
4.2.9 discusses the overall limiting factors in the physical size of the unit based upon structural requirements.
Other limiting factors may be handling operations during stripping, storage, shipping, and erecting. To determine the optimum size of panels and overall economy,
a close collaboration between the designer and precaster is required. Panel sizes, if not integral to the design, can lead to unexpected joint lines, which detract
from aesthetic intentions. The architectural trend has
been to increase both the size and weight of architectural precast concrete panels.
The most economical piece size for a project is usually
a large unit, considering:
1. P roduction repetition and size of available casting
beds.

6. Loads imposed on the support system.
Limitations of dimensions due to handling and storing
vary considerably from plant to plant, but are normally not important considerations for the architect. Tilt
tables (Fig. 3.3.46), strong-backs, stiffening trusses or
pipe frames for handling (Fig. 3.3.47), and plain ingenuity will often allow the larger pieces to be handled.
The precaster may also make use of pretensioning or
post-tensioning to facilitate handling of large units
without risk of cracking or damage.
The architect should bear in mind, however, that
some finishes, such as bushhammering, honing, and
Fig. 3.3.46

Fig. 3.3.47 Temporary strengthening of panels with
significant openings.

Brace

Strongbacks

Tie

Brace

4. A
 vailable crane access and capacity at both the
plant and the project site.

polishing, normally require the panels to be turned
after casting for finishing. Also, the exposed surface
finish may dictate the position of the panel for worker
access during removal of surface retarder or sandblasting. In some cases, it may be necessary to cast in extra
lifting devices to facilitate these maneuvers, and may
add to the cost for larger panels.

5. S torage space, truck turning radius, and other site
restrictions.

Before deeply sculptured elements are designed into
large units, potential storage problems (special racks)

2. H
 andling ease and stability and stresses on the element during handling.
3. T ransportation size and weight regulations and
equipment restrictions.

3

ARCHITECTURAL PRECAST CONCRETE

|

141

3

SURFACE AESTHETICS
3.3.10 Overall Panel Size

should be considered, particularly where the period of
storage may be long or uncertain.
Transportation limitations imposed by product dimensions and weight must be considered during the
design process. It is the precast concrete manufacturer’s option as to which production and transportation
methods will be employed and their responsibility to
verify the behavior of the precast concrete units during these operations. However, the designer should be
familiar with legal highway load limitations and permit
costs that are associated with transporting over-height,
over-width, or over-length members.
Federal, state (provincal), and local regulations may
limit the size, weight, and timing of shipping loads.
Limitations vary from one locale to another whether
the shipment is by truck, barge, or rail. Where climatic
conditions result in load restrictions on some secondary roads during spring thaws, actual timing of the expected delivery should be considered. When large units
are to be moved, a thorough check of local statutes is
mandatory. This is usually done by the shipper, who
should take shipping restrictions into consideration
when planning the route of travel and delivery time to
the site to ensure that sufficient product is delivered in
the prescheduled sequence. This allows for an orderly,
efficient installation in the structure.
The common payload in many areas is 20 ton (18 t)
with a product size restriction of 8 ft (2.4 m) in width,
8 ft (2.4 m) in height, and 45 ft (13.7 m) in length (Fig.
3.3.48). If a unit will fit within these confines, it can
usually be hauled on a standard flatbed trailer without
requiring permits. By use of lowboy (step or drop deck)
trailers, the product height can generally be increased
to about 10 to 12 ft (3.0 to 3.7 m) without requiring
special permits (Fig. 3.3.49). However, lowboys (step
decks) are not as readily available, and their shorter
Fig. 3.3.48 Common overall trucking volume.
53'

8' 0"

45' to

0"
8'

142

|

ARCHITECTURAL PRECAST CONCRETE

bed length may restrict the length of precast concrete
units. A triangular frame mounted on an extendable
lowboy trailer can carry architectural precast concrete
panels as large as 15 × 15 ft (4.5 × 4.5 m) weighing
up to 20,000 lb (9 t) (Fig. 3.3.50). Finger racks allow
architectural precast concrete spandrel beams to be
transported in the vertical position. Note the adjustable screw clamps that secure the beams to the racks
(Fig. 3.3.51).
In most areas, total heights (roadbed to top of product)
of 13 ft 6 in. (4.2 m) are allowed without special permit, while in others this limit is 12 ft (3.7 m). On occasion, even such heights require special routing to avoid
low overpasses and overhead restrictions. Restrictions
generally exist for loads over 8 ft (2.4 m) in width; maximum permit widths can vary from 10 to 14 ft (3.0 to
4.3 m) depending on the area or city. Some areas allow
overall lengths over 70 ft (21 m) with only a simple permit, front and rear escorts, and travel limited to certain
times of the day. Beside variations in lengths, heights,
and widths, weight restrictions vary widely. Thus, the
general load limit without permit (20 to 22 ton) (18
to 20 t) can, in some areas, be increased to 100 ton
with special permit, while in others there are very severe restrictions on loads over 25 ton (23 t). Permits are
issued in most areas only for non-divisible loads. These
restrictions will add to the cost of the precast concrete
unit, and should be evaluated against savings realized
by combining smaller units into one large unit.
In determining final dimensions, consideration should
be given to utilizing a full truckload. A precast concrete
unit or several units should approximate this usual payload. For example, an 11-ton (10 t) unit is not economical, because only one unit can be shipped per load,
whereas a 10-ton (9 t) unit would be economical, because two units per load can be shipped. For quick calculations, normalweight concrete, including reinforcement and hardware, weighs 150 lb/ft3 (2400 kg/m3).
In order to facilitate erection, it is desirable to transport members in the same orientation they will be on
the structure. In many cases this is possible. For example, with single-story wall panels, transportation
can be accomplished on an A-frame type trailer with
the panels in an upright position from which they can
be lifted directly into position. With this type of trailer,
good lateral support, as well as two points of vertical
support, are provided to the members (Fig. 3.3.52[a]).
Longer units, which are thin compared to their length

SURFACE AESTHETICS

3.3.10 Overall Panel Size

3

Fig. 3.3.49

Fig. 3.3.51

Fig. 3.3.50

Fig. 3.3.52(a)

and width, can be transported in a favorable orientation to reduce tensile stresses. Two- or three-story
panels can be transported on their long sides, taking advantage of increased stiffness while supporting
the panel on two or more points, with lateral support
along the length of the panel (Fig. 3.3.52[b]).
The cost of transportation is not proportional to the
distance covered as the cost of loading and unloading
(plus protection of the load) becomes less significant
on longer hauls. The rate per ton mile is normally reduced for longer hauls. Consequently, long hauls have

Fig. 3.3.52(b)

occurred on competitively bid jobs.
Piggyback transport by rail has been used successfully, but otherwise rail transportation is not common
because of the delicate coordination required and the
potential damage to units.
Water transportation via barge is inexpensive and
safe, but even for plants with navigable water frontage, double handling will occur when the barges reach
their destination because the panels will normally have
to be transferred to the site by trucks.

ARCHITECTURAL PRECAST CONCRETE

|

143

3

SURFACE AESTHETICS
3.4 Colors and Textures

3.4 C OLORS AND TEXTURES
3.4.1 C olors
Architectural precast concrete can be cast in almost
any color, form, or texture to meet aesthetic and functional requirements of the designer in an economical
manner. Complementary combinations of color and
texture can aesthetically improve any project.
Design flexibility is possible in both color and texture of
precast concrete by varying aggregate and matrix color,
size of aggregates, finishing processes, and depth of
aggregate exposure. Combining color with texture accentuates the natural beauty of aggregates (Fig. 3.4.1).
Aggregate colors range from white to pastel to red,
black, and green. Natural gravels provide a wide range
of rich warm earth colors, as well as shades of gray.
Specifying color and texture in precast concrete is
not a difficult, laborious, or seemingly impossible task.
Fortunately, there is a resource available to the specifier
that can make this task easy when the specifier does
not have a sample to match. It is the PCI Architectural
Precast Concrete – Color and Texture Selection Guide,
2nd Edition. The guide was specifically developed as
a starting point for the selection of color and texture.
It contains several hundred images of colors and textures, and their associated mixture materials, that can
be achieved with architectural precast concrete.
PCI’s Color and Texture Selection Guide is available
for viewing at www.pci.org. The online guide illustrates the world of possibilities of architectural precast
concrete options in color and texture.

Color and, consequently, color tone represent relative
values. They are not absolute and constant, but are affected by light and shadow, intensity, time, and other
surrounding or nearby light-reflecting colors. A concrete surface, for instance, with deep-exposed opaque
white quartz appears slightly gray. This is due to the
shadows between the particles blending with the actual color of the aggregate and producing the graying
effect. These shadows in turn affect the color tone of
the matrix.
Similarly, a smooth concrete surface will change in
tone when striated. Also a white precast concrete window unit with deep mullions will change tone when
bronze-colored glass is installed. Color tone is constantly changing as the sun traverses the sky. A clear
sky or one that is overcast will make a difference, as
will landscaping and time. And last, but by no means
least, in large city and industrial environments, air pollution can cause color tone to change.
Color selection should be made under lighting conditions similar to those where the precast concrete will
be used, such as the strong light and shadows of natural daylight. Muted colors usually look best in subdued
northern light. In climates with strong sunlight, much
stronger and brighter colors are used with success.
Surface texture also affects color. A matte finish will
result in a different panel color than does a smooth finish. Texture helps to determine the visual importance of
a wall and, hence, the color. For example, moderately
rough finishes usually are less obtrusive than shiny surfaces. The building’s appearance is a function of the
designer’s use of light, shadow, texture, and color.
Matrix color (cement plus pigment) exerts the primary
color influence on a smooth finish because it coats the
exposed concrete surface. As the concrete surface is
progressively removed and the aggregates are exposed,
the panel color increasingly reflects the fine and coarse
aggregate colors. Nevertheless, the matrix color always
has an effect on the general tone of the panel.

Fig. 3.4.1

144

|

ARCHITECTURAL PRECAST CONCRETE

Cement may be gray, white, or a mixture of the two.
All cements possess inherent color and shading differences depending on their brand, type, mill, and quarry
source. For example, some gray cements are nearly
white while others have bluish, reddish, or greenish
tones. Some white cements have a buff or cream undertone, while others have a blue or green undertone.
In addition, a finely ground gray or white cement is

SURFACE AESTHETICS

3.4.1 Colors

normally lighter in color than a coarse ground cement
of the same chemical composition. If color uniformity
is essential, cements of the same type and brand from
the same source should be specified. Gray cement is
generally subject to greater color variation than white
cement even when supplied from one source. Normal
production variables, such as changes in water content, curing cycles, temperatures, humidity, and exposure to climatic conditions at varying strength levels, all
tend to cause greater color variations in a gray cement
concrete in relation to concrete made with white cement. A low water-cement ratio cement paste is almost always darker than a high water-cement ratio
paste made with the same cement.
Although white cement will give the least amount
of color variation, it is important to choose the lightest color aggregates to decrease the shadowing effect
of aggregates close to the surface. Gray cement has
a greater ability to provide an opaque covering of aggregate, but has color differences that may offset this
advantage.
Different combinations of gray cement, white cement,
pigments, and aggregates offer an extensive range of
possible color combinations. If gray is the desired color
of the matrix and the optimum uniformity is essential,
a mixture of white and gray or white cement with gray
pigment is recommended. Uniformity normally increases with increasing percentages of white cement,
but the gray color remains dominant. See Section 2.2.5
for economy of materials.
Pigments often are added to obtain the desired matrix color. All pigments used should pass ASTM C 979,
Pigments for Integrally Colored Concrete. Most pigments used to color concrete are iron oxide pigments—
both natural and synthetic—and hence tend to be
earth tones. Natural iron oxides are widely available in
earth tone red, yellow ochres, and raw and burnt umbers. Synthetic iron oxides are manufactured in shades
of red, yellow, and black. Other pigments are available
to achieve green and blue shades.
Different amounts of pigment, expressed as a percentage of the cement content by weight, produce
various shades of color. All pigments, however, have
a saturation point, beyond which additional pigment
will not continue to significantly increase the color intensity. For synthetic iron oxides this saturation point
is around 5% and for natural iron oxides it is around

3

10%. Pigment additions should never exceed 10% of
the weight of the cement in the mixture. High percentages of pigment reduce concrete strength because of
the high percentage of fines introduced into the mixture by the pigments (which increases the water requirement of the mixture). For this reason, the amount
of pigment should be controlled within the limits of
strength and absorption requirements.
Pigments may be combined to achieve the desired
shade, as long as the total amount stays below the
recommended maximum level. White portland cement must be used to create light pastel shades such
as buff, cream, ivory, pale pink, and rose tones, as well
as bright intense yellows, oranges, and reds. Red, tan,
black, dark gray, and other hues are produced very satisfactorily using gray cement.
Shades of buff, tan, red, orange, yellow, brown,
charcoal, and gray are the least costly. Green is permanent but expensive, except in light shades. Blue is
very expensive and some blues are not uniform or permanent. Cobalt blue should be used to avoid permanence problems. Dark black colors cannot be achieved
in concrete. The use of carbon black can initially create
intense black concrete. However, due to its extremely
fine particle size, it has a tendency to leach out of the
concrete matrix. As the pigment is removed, the concrete substrate appears increasingly “faded.” Synthetic
black iron oxide, on the other hand, will produce a stable charcoal color.
Titanium dioxide pigment, in addition rates of 1 to
3% of the weight of the cement, is sometimes used to
lighten gray concrete or to further intensify the whiteness of white concrete. However, titanium dioxide does
not have the tinting power to produce white concrete
when using gray cement. If white concrete is desired,
white cement must be used.
Many of these pigment options are represented in Fig.
3.4.2. Each column shows variations in cement color:
gray, 1/2 gray and 1/2 white, and all-white portland cement. Each grouping of three samples represents a
different pigment shown at three different concentrations. The amount of pigment added is expressed as
a percentage by weight of the cement in the mixture.
These swatches are meant to represent a range of possibilities for matrix colors available in architectural precast concrete.

ARCHITECTURAL PRECAST CONCRETE

|

145

3

SURFACE AESTHETICS
3.4.1 Colors

Fig. 3.4.2 Color selections.

146

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.4.1 Colors

ARCHITECTURAL PRECAST CONCRETE

|

3

147

3

SURFACE AESTHETICS
3.4.1 Colors

Significant points to consider when color consistency
is critical are:
1. Quality and quantity of the pigment.
2. Proper batching and mixing techniques and the
coloring agent’s affect on concrete workability.
3. Quality (freedom from impurities) of the fine and
coarse aggregates.
4. Uniform quantities and gradation of fine materials
(passing No. 50 sieve and including the cement) in
the concrete mixture.
5. Careful attention to curing and uniform duplication of curing cycles.
6. Type and color of cement.
7. Constant water-cement ratio in the mixture.
8. Consideration of those factors that can contribute to efflorescence. This is especially important
for dark and intense colors, which aggravate any
efflorescence problem by making it more visible.
Efflorescence deposited on the surface may mask
the true color and give the appearance of pigment
fading even though the pigment or cement paste
has undergone no change. The original color may
be restored by washing with a dilute solution of hydrochloric acid and water and rinsing thoroughly.
In addition, weathering of the pigmented cement
paste exposes more of the aggregate to view. If the
color of the aggregate contrasts that of the pigment, a change in the overall color of the surface
may be noted.
Fine aggregates (sand) have a major effect on the
color of white and light-colored concrete, and can add
color tones. Where the color depends primarily on the
fine aggregates, gradation control is required, particularly where the color tone depends on the finer particles. Where fine aggregates (sand) are manufactured
by crushing colored coarse aggregates and bagged by
sizes directly from the screening operations, uniformity in gradation can be maintained from one batch
to the next. For fine aggregates in bulk, that are subject to several rehandling processes, this is not feasible.
Consequently, it is recommended that for bulk material, the percentage of the fine aggregates passing
the No. 100 (150 µm) sieve should be limited to no
more than 5%. This may require washing of such aggregates, but the premium for this is justified by the
increased uniformity of color.
The precast concrete manufacturer should verify

148

|

ARCHITECTURAL PRECAST CONCRETE

that adequate supply from one source (pit or quarry)
for each type of aggregate for the entire job will be
readily available. If possible, the precaster should obtain the entire aggregate supply prior to starting the
project, or have the aggregate supply held by the aggregate supplier. Stockpiling will minimize color variation caused by variability of material and will maximize
color uniformity.
For reasons of workability, a percentage of natural
sand is preferable in a concrete mixture. Manufactured
sand, however, often adds valuable color tones, and may
be used as part of the fine aggregates. Manufactured
sand is generally more expensive than natural sand and
may not always be available. Crushed pink granite, for
instance, will create a warm-colored matrix, but due
to the size of sand particles, the pink can hardly be
distinguished in the finished unit. With a light to medium exposure, a uniform color appearance may be
obtained by using crushed sand of the same material
as the coarse aggregate. When maximum whiteness
is desired, a natural or manufactured, opaque white
or light yellow sand should be used. Most naturally
occurring sands lack the required whiteness, and the
precaster usually must look to the various manufactured fine aggregates to achieve the white color desired. Generally, these aggregates consist of crushed
limestone (dolomite, calcite) or quartz and quartzite
sands.
The colors in coarse aggregate are multiple, and most
precasters will have a supply of aggregate samples at
the plant. Selection of aggregates for colors should be
governed by the following considerations:
1. Aggregates must measure up to proper durability (soundness and absorption) requirements, be
free of impurities (iron oxides), and be available in
shapes required for good concrete and appearance
(chunks rather than slivers).
2. Aggregate shape affects the appearance of a surface after weathering. Rounded aggregates (pebbles) tend to remain clean, but angular aggregates
of rough texture tend to collect dirt, and confine
it to the recesses of the matrix. For this reason, as
well as architectural appearance, the area of exposed matrix between aggregate particles should
be minimized. It may be advisable for the matrix to
be darker than the aggregate in structures subject
to considerable atmospheric pollution.
3. Final selection of colors should be made from

SURFACE AESTHETICS

3.4.1 Colors

concrete samples that have the proper matrix and
are finished similarly to the planned production
techniques. Some finishing processes change the
appearance of aggregates. Sandblasting will dull
them, while acid etching may increase their brightness. Exposure by retardation normally leaves the
aggregates unchanged. The method of exposing
aggregate alters the color of the surface by affecting the color of the aggregate and by the amount
of shadow cast by the exposed particles.
4. Aggregates with a dull appearance in a gray matrix
may well appear brighter where the matrix is basically light colored.
5. Weathering may influence newly crushed aggregates. When first crushed, many aggregates are
bright but may dull slightly with time. Similarly,
some of the sparkle caused by acid etching or
bushhammering may not survive more than a few
weeks. The architect should recognize that samples maintained indoors may not retain their exact
appearance after exposure to weather even after
a few weeks.

3

Coarse aggregates are selected on the basis of color,
hardness, size, shape, gradation, method of surface
exposure, durability, cost, and availability. Colors of
natural aggregate vary considerably according to their
geological classification and even among rocks of one
type (Fig. 3.4.3).
Clear quartz provides a sparkling surface to complement the color effect created by use of a pigmented
matrix. White quartz ranges from translucent white
to a deep milky white. Rose quartz provides surfaces
ranging from a light pink to a warm rose color. Green,
yellow and gray colors are also available.
Marble offers the widest selection of colors, including
green, yellow, red, pink, blue, gray, black, and white.
Blue and yellow marble aggregates are available in pastel hues. Marble is available in many shades running
from light to moderately dark. Crushed limestones
tend toward white, gray, and pink colors.
Granite in shades of pink, red, gray, dark blue, black,
and white produces a soft, mottled appearance when
used in concrete. Traprocks, such as basalt, provide gray,
black, and green colors and are particularly durable.

Fig. 3.4.3 Kaleidoscope of aggregate colors.

Photos:Wyckoff Advertising, Inc.

ARCHITECTURAL PRECAST CONCRETE

|

149

3

SURFACE AESTHETICS
3.4.1 Colors / 3.4.2 Textures

Some washed and screened gravels can be used
to provide brown or reddish-brown finishes. Yellow
ochers, umbers, and sandy (buff) shades abound in
river bed gravels. Also, an almost pure-white gravel occurs in several sedimentary formations.
Marine-dredged aggregates or seashells should be
washed with fresh water to reduce their salt content.
There is no maximum limit on the salt content of coarse
or fine aggregate; however, the chloride limits for the
concrete should be followed. Seashells are hard materials that can produce high-quality concrete. Due to
the angularity of the shells, additional cement paste is
required to obtain the desired workability. Aggregate
containing complete shells (uncrushed) should be
avoided, as its presence will result in voids in the concrete and lower the compressive strength.
Local aggregates should not be overlooked. They usually are economical and can be attractive with the proper
matrix. Local architectural precasters are familiar with
available aggregates and usually have concrete samples
made with different materials on display (Fig. 3.4.1).
Coarse aggregates should be reasonably uniform in
color. However, surfaces consisting of a single color
lack clarity and, strangely enough, purity. In general, a
light-colored aggregate is preferable to avoid shaded or
toned areas. Light and dark coarse aggregates require
care in blending to provide color uniformity within a
single unit. With a small color difference between the
light and dark aggregates and a small variance in total
amounts of each aggregate, the chances of uniformity
are enhanced. It is advisable to match the color or tone
of matrix to that of the coarse aggregate so that minor
segregation of the aggregate will not be noticeable.
Panels containing aggregates and matrices of contrasting colors will appear less uniform than those containing materials of similar colors (as the size of the coarse
aggregate decreases, less matrix is seen and the more
uniform the color of the panel will appear).
The choice of aggregates becomes more critical in
smooth white concrete. Due to the greater difference
in color between the white cement and aggregates,
the white cement has less ability than gray cement to
form an opaque film over the aggregates and prevent
the aggregate color from showing through. Thus, special consideration must be given to the selection of
suitable aggregates to help prevent variations in color
and color intensity on the finished surface. A light-colored aggregate is preferable to a dark aggregate when

150

|

ARCHITECTURAL PRECAST CONCRETE

trying to avoid shaded or toned areas.
Two concrete mixtures with differently colored matrices exposed at the face of the same panel should
be specified only with a demarcation feature such as a
rustication or protruding area.
The ease of obtaining uniformity in color is directly related to the ingredients supplying the color. Optimum
uniformity is obtained by using white cement. Extreme
color differences between aggregates and matrix
should be avoided. In all cases, color should be judged
from a full-sized sample that has the proper matrix and
has been finished in accordance with planned production techniques.

3.4.2 Textures
Textures allow the naturalness of the concrete ingredients to be expressed, provide some scale to the mass,
express the plasticity of the concrete, and improve its
weathering characteristics. A wide variety of textures is
possible, ranging from a honed or polished surface to
a deeply exposed one.
The surface finish enhances the character of the
building by contributing a presence to complement the
building aesthetics. However, a small, solitary concrete
sample can mislead the architect in the value of a finish
compared to its appearance when viewed in the building scale from a distance.
As a general rule, a textured surface is more aesthetically pleasing (greater apparent uniformity) than
a smooth as-cast finish. The surface highlights and the
shadings of aggregate color camouflage subtle differences in texture and color of the concrete. Also, any
damage is more easily repaired on textured surfaces
than on smooth finishes.
A texture may be defined, in comparison with a
smooth surface, as an overall surface pattern. The
range of textured finishes for architectural precast concrete molds includes the characteristic imprint or patterns created from a form liner or mold. Alternative
textured finishes may be produced by removing the
surface mortar to expose the coarse aggregate in the
mixture by various methods.
A profile may be defined, in comparison with a flat
surface, as a shape rather than a texture, produced
from a specially made mold or form liner. One well
known example is the striated or ribbed finish.

SURFACE AESTHETICS

3.4.2 Textures

Profiled surfaces can be either smooth or textured,
in a similar way flat surfaces can be either smooth or
textured. This gives four possible combinations.

3

Appearance at 75 ft (23 m)

It is also possible for part of a panel to be given more
than one finish. This design feature allows for a wide
range of appearance options. A detailed description
of the more common textures and finishes is given in
Section 3.5.
There are four important factors to be considered in
choosing a texture:
1. The area of the surface. This affects the scale
of the texture. Coarse textures usually cannot be
used effectively for small areas. Dividing large, flat
areas or surfaces into smaller ones by means of
rustications tends to deemphasize any variations in
textures.

Fig. 3.4.4

2. The desired effect at a viewing distance. The
designer may seek a visually pronounced texture or
may use texture as a means to achieve a particular
tone value. The visual effect desired at the normal
viewing distance influences the texture and size of
aggregate chosen for the panel face. Figure 3.4.4
shows different size aggregates viewed at 30 ft (9
m) and 75 ft (23 m).

Table 3.4.1 Suggested Visibility Scale.

Suggested visibility scale
Aggregate size,
in. (mm)
1
/4–1/2 (6–13)
1
/2–1 (13–25)
1–2 (25–50)
2–3 (50–75)

Distance at which texture is visible, ft (m)
20–30 (6–9)
30–75 (9–23)
75–125 (23–38)
125–175 (38–53)

The viewing distances in Table 3.4.1 are based
on the use of aggregate of one color. They may
require modifications when the aggregate contains
both light and dark particles. Further modifications
may be required to include the effects of panel
orientation. For example, the contrast caused by
shadows from aggregate particles will vary with
lighting conditions.
3. The orientation of the building wall elevation.
This determines the amount and direction of light
on the surface and how the panel will weather.

Appearance at 30 ft (9 m)

4. Aggregate particle shape and surface characteristics. For exposed-aggregate textures, the aggregate particles may be rounded, irregular, angular, or flat. Their surfaces may be glossy, smooth,
granular, crystalline, pitted, or porous. Both the
shape and surface characteristics determine how
the surface will weather and reflect light.
In addition to the visual effect of texture within reasonable distances, textures may be used to achieve
colors based on the natural colors of the exposed aggregates and matrix.
The size of the aggregate should be related to the
configuration of the panels. The larger the aggregate,
the more difficult it will be to detail edges, reveals, and
returns.

ARCHITECTURAL PRECAST CONCRETE

|

151

3

SURFACE AESTHETICS
3.4.2 Textures

Exposed-aggregate finishes are popular because
they are reasonable in cost and provide an infinite
variety of colors and textures. This is achieved by
varying the type, color, and size of aggregate, color of matrix, method of exposure, and depth of
exposure.
The different degrees of exposure are:
L ight Exposure — where the surface skin of
cement and some sand is removed, just sufficiently to expose the edges of the closet
coarse aggregate. This imparts a fine, sandy
texture. Matrix color will greatly influence the
overall panel color.
	Medium Exposure — where further removal
of cement and sand has caused the coarse aggregate to visually appear approximately equal
in area to the matrix.
	Deep Exposure — where cement and sand
have been removed from the surface so that
the coarse aggregate becomes the major surface feature.

Sandblasted

The extent aggregates are exposed or “revealed”
is largely determined by their size. Exposure should
not be greater than one-third the average diameter
of the coarse aggregate particles or one-half the
diameter of the smallest sized coarse aggregate.
Fig. 3.4.5 Different degrees of exposure.

Acid Etched
Retarded

Figure 3.4.5 shows photographs of retarded,
sandblasted, and acid-etched samples with the
various depths of exposure on the same concrete
mixture.

152

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5 Finishes / 3.5.2 Smooth As-Cast

3.5 F INISHES
3.5.1 G
 eneral
Finishes, in terms of color and textures, are discussed
in Section 3.4. This section describes the various methods of obtaining these finishes. Because surface finishes depend on properly fabricated molds, the designer
should clearly understand the capabilities and limitations of mold production (see Section 2.2).
The appropriate finish should be carefully chosen and
clearly specified. The designer should base the final
choice of surface finish on a balance between appearance (uniformity of color and texture) and cost with
consideration of the limitations in materials and production techniques. The appearance can be judged
using a combination of samples and reduced-scale or
full-scale mockups (see Section 3.2). These samples
or mockup panels can then be made available at the
precast concrete plant so that all concerned can be assured that standards of finish and exposure are being
maintained. Appearance, colors, and textures of the
surface finishes of all units should match within the
acceptable range of the colors, textures, and general
appearances of the approved sample panels.
During the manufacturing process different panels
may be subjected to varying levels of ambient humidity. Initially, tonal variations in color might be considered unsatisfactory, but are likely to moderate when
the panels have a balanced moisture content.
Quality assessment should also include the likelihood
of maintaining a reasonable level of uniformity from
start to finish of production. For instance, it is not too
difficult to get a uniform distribution of two differently
colored aggregates in a small sample produced under
laboratory conditions, but it could be a difficult task to
produce the same uniform appearance on a daily basis.
Generally speaking, if two different colored aggregates
are contemplated, the difference in appearance (colors)
should not be too prominent, and similarly, the color
difference between aggregates and matrix should also
be weighed against the practicality of obtaining a uniform appearance.
A compromise may be required between the finish
and the shape of a precast concrete panel. Sculptured
panels may have visible air voids on the returns that
become accentuated when the surface is lightly finished. Normally, smooth finishes also will have air voids
on return surfaces. If air holes are of a reasonable size

3

(1/8 to 1/4 in. [3 to 6 mm]), it is recommended that they
be accepted as part of the texture. Filling and sackrubbing will eliminate the voids, but this method is
expensive and may cause color differences. Exposedaggregate finishes often have variations between faces
and returns. To minimize differences, mixtures should
contain reasonably spherical or cubical aggregates.
For large returns, or other situations where variations
in appearance must be minimized, sequential casting
should be considered.
The surface of large panels should be divided into
smaller areas by means of rustications or reveals to
minimize the perception of textural differences.
Finishing techniques used in individual plants may
vary considerably from one part of the continent to
another, and between individual plants. Each plant
has developed specific finishing techniques supported
by skilled operators and/or special facilities. For actual
projects, be sure to confer with the local precasters for
assistance in obtaining the desired appearance and
relative costs. The following sections discuss each of
the finishing techniques.

3.5.2 Smooth As-Cast
A smooth as-cast finish shows the natural look of the
concrete without trying to simulate any other building
product. Fine surface details and sharp arrises can be
achieved with a smooth finish. This finish is perhaps the
most difficult to produce. When a high level of color
uniformity is required, its use is strongly discouraged.
There is also the question of how the surface will change
when exposed to the weather (see Section 3.6). Smooth
surfaces tend to weather unevenly and become discolored from rainwater and airborne particles.
Smooth concrete makes the maximum demands
on the quality and maintenance of the mold and on
the concrete itself. Color variations tend to be most
pronounced when the mold face is glassy and impermeable. While a rough concrete surface will scatter
reflected light and soften the impact of blemishes, a
smooth surface will make variations more conspicuous.
Color uniformity is difficult to achieve on gray, buff,
and pigmented concrete surfaces. The use of white cement yields better color uniformity than gray cement.
Allowable color variation in the gray cement is readily apparent on the uninterrupted surfaces of smooth
off-the-mold concrete, and any variation is likely to be
regarded as a surface blemish.

ARCHITECTURAL PRECAST CONCRETE

|

153

3

SURFACE AESTHETICS
3.5.2 Smooth As-Cast

(a)

Fig. 3.5.1(a), (b) & (c)
Prince of Peace Catholic Church
Taylors, South Carolina;
Architect: Craig Gaulden Davis Inc.
(b)

The core of the church in Fig. 3.5.1(a), (b) and (c) is a
loadbearing precast concrete frame that also serves as
the primary interior and exterior finish. Each exposed
precast concrete element is composed of two pieces,
joined back to back. They are stacked and joined to elements above and below with steel pins. The smoothas-cast gray surfaces serve as final finishes and allow
the church environs to recreate a historic way of building by rendering the appearance of stone. The detailing and surface treatment of the precast concrete
components satisfied aesthetic concerns and the use
of the structural material as the building’s primary finish lent integrity to the design concept.
Conceived as a work of art in its own right, the art
center in Fig. 3.5.2(a) was designed to interact with its
surroundings and dispense with the traditional walls
of a museum. The main façade is restrained with horizontal windowless volumes of gray, smooth as-cast
precast concrete and black-anodized aluminum panels
effortlessly floating over the glass voids. The architect
achieved the desired raw as-cast look with precast
concrete. At the corner of the building, the cantilevered projections of the block-like elements become
more defined (Fig. 3.5.2[b]). The vertically stacked
masses seem like a cubist collage of a façade hovering
above the glazed base. The panels appear to almost

154

|

ARCHITECTURAL PRECAST CONCRETE

(c)

defy gravity while approximating the cacophony of the
streetscape.
Smooth as-cast precast concrete panels have a
smooth film of hardened cement matrix. The finished
color is therefore determined primarily by the color of
the cement. In some instances the sand may also have
some affect. Initially, this is likely to be insignificant unless the sand contains a high percentage of fines or is
highly colored. However, as the surface weathers, the
sand becomes more exposed and its influence or effect
on color becomes more marked.
The color of the coarse aggregate should not be significant unless the particular panel requires a high degree of consolidation. Under this circumstance, some
aggregate transparency may occur, causing a blotchy,
non-uniform appearance. Aggregate transparency, or
“shadowing,” is a condition in which a light-colored,
formed concrete surface is marked by dark areas simi-

SURFACE AESTHETICS

3.5.2 Smooth As-Cast

lar in size and shape of particles of coarse aggregate
in the concrete mixture. When encountered, it usually
appears on smooth surfaces and causes a blotchy and
irregular appearance. The effect can be reduced by using lightly colored coarse aggregates with low absorption and white cement.

to 1/4 in. (3 to 6 mm), it is recommended that they
be retained as part of the surface texture rather than
sack-rubbed. Filling and sack-rubbing will eliminate
the voids, but many cause increased color variations.
Samples or mockup panels should be used to establish
the acceptability of color variation and air voids with
respect to frequency, size, and distribution uniformity.

The smooth cement film on the concrete may be susceptible to surface crazing (fine and random hairline
cracks) when exposed to wetting and drying. In most
cases, this is a surface phenomenon (penetrates only
as deep as the thin layer of cement paste at the surface of a panel) and does not affect structural properties or durability. In some environments, crazing will be
accentuated by dirt collecting in these minute cracks.
This will be more apparent in white than gray finishes
and on horizontal more than vertical surfaces.

For true economy, units with smooth as-cast surfaces
should be produced without additional surface treatment after stripping from the molds, except for possible washing and cleaning. This, in turn, demands the
following precautions:
1. Provide architectural relief to flat, exposed surfaces.
Some sculpturing of the panel is highly desirable.
Careful attention to detailing is essential. Make
provisions for ample draft, chamfer edges and corners to minimize stripping damage (see Sections
3.3.2 and 3.3.7), and provide suitable water drips
and other weathering details (see Section 3.6). The
smooth as-cast finishes are the most difficult of all

Precast concrete panels with a smooth as-cast finish will normally have air voids, particularly on return
surfaces. If these air holes are of reasonable size, 1/8
(a)

3

(b)

Fig. 3.5.2(a) & (b)
Lois & Richard Rosenthal Center for Contemporary Art
Cincinnati, Ohio;
Architect: Zaha Hadid Architects, Design Architect;
KZF Design Inc., Architect of Record.

ARCHITECTURAL PRECAST CONCRETE

|

155

3

SURFACE AESTHETICS

3.5.2 Smooth As-Cast / 3.5.3 Exposed Aggregate by Chemical Retarders and Water Washing

precast concrete finishes to repair in terms of color
and texture match.
2. The architect should specify the panel surface expectations regarding joints in the mold face, so
that the acceptable level of surface appearance is
established.
The designer and precaster must accept and understand the limitations of smooth as-cast finishes.
Smooth, as-cast precast concrete panels usually have
some surface imperfections. Minor variations in texture of mold surface reflected on the smooth concrete
surface, color variations, air voids (bug/blow holes),
and minor surface crazing and blotchiness are to be
expected, especially on non-profiled flat panels. Both
designer and precaster must be aware of the realistic surface finish that will be obtained. Of all precast
concrete finishes, this finish is the most misunderstood

Fig. 3.5.3
Commissioners of Public Works Administrative Offices
Charleston, South Carolina;
Architect: Lucas Stubbs Pascullis
Powell & Penney, Ltd.;
Photo: LS3P Associates Ltd.

when it comes to acceptability. An acceptable smooth
finish can be very difficult and expensive to achieve if
a high degree of uniformity is anticipated by the architect or owner. If the surface is to be painted or stained,
this finish will provide an excellent surface, while keeping costs to a minimum.
Many of the aesthetic limitations of smooth concrete
may be minimized by the shadowing and depth provided by profiled surfaces (fluted, sculptured, board
finishes, etc.), by subdividing the panels into smaller
surface areas by means of vertical and horizontal reveals or rustications, or by using white cement (Fig.
3.5.3). Any introduction of shapes to provide shadow
effects will enhance the final finish.

3.5.3 Exposed Aggregate by Chemical
Retarders and Water Washing
Chemical surface retarders provide a non-abrasive
process that is very effective in bringing out the natural
color and luster of coarse aggregate. The application
of a chemical retarder to the mold surface prior to casting the concrete delays the surface cement paste from
hardening within a time period and to a depth dependent on the type or concentration of retarder. After
hardening of the concrete mass (normally overnight),
the retarded outer layer of cement paste is removed by
high-pressure water washing, exposing the aggregate
to the desired depth. This process should take place at
a predetermined time after casting (Fig. 3.5.4).
The aggregate exposure obtained is controlled by the
retarder; therefore, any variations in exposure are not
as correctable as with sandblasting. Furthermore, deep
retarded surfaces do not allow for sharp panel profiles.
The shape of the coarse aggregate, its position after
consolidation, and the depth of exposure will determine the surface appearance. Appearance will therefore vary to some degree with surface orientation and
aggregate shape. This is especially critical in units with
returns, where vertical sides are expected to reasonably match the bottom face (see Section 3.3.7).
Precast concrete cladding matched the existing buildings’ limestone and ashlar-laid red granite masonry,
and successfully produced compatibility with the campus’ collegiate gothic architecture (Fig. 3.5.5[a]). The
red granite appearance was produced using a blend
of gray and white cements, red pigments, granite aggregate, and a retarder and washing to expose the ag-

156

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.3 Exposed Aggregate by Chemical Retarders and Water Washing

3

and returns (see Section 3.3.8).
If the bright, natural colors of the aggregate are the
prime concern, exposed aggregate from retarded surfaces is the best way to achieve this result. The mixture
proportions, aggregate gradation and physical characteristics of the aggregate, and matrix/aggregate color
compatibility are important. It is advisable to vary the
color or tone of the matrix wherever possible to match
or blend in with the color of the aggregate. This match
can be achieved by careful selection of cement, sand,
and pigment colors. A good matrix to coarse aggregate
color match will minimize mottled effects (minor variation in aggregate distribution) from being noticeable.

Fig. 3.5.4 Removal of retarded surface.

gregate. Quarry cut limestone was simulated using the
same cements, limestone aggregate, pigment, and a
light sandblast finish (Fig. 3.5.5[b]).
Depending on the particular mold configuration (vertical, radius, or complicated shapes), the placing of
concrete may scour the retarder applied to sloping surfaces and affect the finish of the concrete. These factors may compound the problem of matching bottoms

Chemical retarders are also available for the face-up
method of casting concrete. Retarder is sprayed on the
concrete surface following consolidation and finishing operations. Because the consolidation of concrete
brings an excess of mortar and water to the surface,
an exposure of this surface will normally fail to show as
dense a coarse aggregate distribution as a down-face
mold surface. This may be overcome by seeding the
(b)

Fig. 3.5.5(a) & (b)
Washington University Hilltop Campus Parking
St. Louis, Missouri;
Architect: Jacobs Technologies (formerly Sverdrup Facilities);
Skidmore, Owings & Merrill, Design Consultant;
Photos: Max Rogers.
(a)

ARCHITECTURAL PRECAST CONCRETE

|

157

3

SURFACE AESTHETICS

3.5.3 Exposed Aggregate by Chemical Retarders and Water Washing

(b)
Fig. 3.5.6(a) & (b)
One Boca Place
Baca Raton, Florida;
Architect: Smallwood, Reynolds, Stewart,
Stewart & Associates;
Photos: Gabriel Benzur.

(a)

(b)
Fig. 3.5.7(a) & (b)
Adtran Corporate
Headquarters, Phase IV
Huntsville, Alabama;
Architect: Cooper Carry Inc.;
Photos: (a) Gabriel Benzur;
(b) Steve Brock.

(a)

158

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.3 Exposed Aggregate by Chemical Retarders and Water Washing

surface with the coarse aggregates following the initial
finishing of the surface, then tamping or rolling the
aggregates into the surface, refinishing the surface,
and applying the retarder. The spandrels and sunscreen
panels shown in Fig. 3.5.6(a) and (b) are exposed on
all sides. Uniformity of face-up exposed aggregate will
not match the face-down surface and as such its use
should be minimized.

3

Fig. 3.5.9
Rhode Island Convention Center – North Parking Structure
Providence, Rhode Island;
Architecture: Cannon;
Photo: Lucy Chen.

The appearance of aggregate in precast concrete
units subjected to retardation will not change from the
natural appearance of these aggregates prior to incorporation in the concrete mixture. These methods may
be used for all three degrees of exposure, but they are
most commonly used for medium or deep exposure.
Figures 3.5.7(a) and (b) show a retarded finish simulating red, flamed granite. Retarded and water-washed
finishes are relatively easy to repair, a major advantage.
Also, the mold surface is not as critical if the aggregate is to be exposed. Figure 3.5.8 shows the different
appearances resulting from a medium retarded finish
(top) and a medium sandblast finish (bottom) with the
same concrete mixture design. The exterior of the parking structure (Fig. 3.5.9) is sheathed in precast concrete
panels with medium sandblasted and exposed (retarded), rose-colored–granite aggregate finishes. The
structure in Fig. 3.5.10 also has retarded and abrasive
blasted finishes produced from the same mixture.
Demarcation or rustication features are recommended to prevent ragged edges where retarder is applied
to a part of a mold surface.

Fig. 3.5.8

Fig. 3.5.10 Retarded and abrasive blasted finishes produced
from the same concrete mix.

ARCHITECTURAL PRECAST CONCRETE

|

159

3

SURFACE AESTHETICS
3.5.4 Form Liners and Lettering

3.5.4 F orm Liners and Lettering
An almost unlimited variety of attractive patterns,
shapes, and surface textures can be achieved by casting against wood, steel, plaster, elastomeric, plastic,
or polystyrene-foam form liners (Fig. 3.5.11). These
form liners may be incorporated in or attached to the
surface of a mold. Concrete’s plasticity offers the opportunity for innovation and individual character in the
surface textures, patterns, and shapes, which can be
achieved by casting against the various types of form
liners. A large pattern offers ever-changing details due
to the play of light and shadow; a fine texture offers
a muted appearance that is subtle but not drab and
smooth surfaces bring out the elegance and richness
of simplicity. Form liner textured surfaces also mask
minor imperfections that would otherwise be obvious
in a smooth as-cast surface, yielding a more uniform
appearance.
Light and shade created by modeling or sculpturing with liners may be used for visual effect to enliven
large concrete surfaces with low relief patterns at a
reasonable cost or can economically simulate another
material in concrete.

Fig. 3.5.11 Some of the available form liner patterns.

160

|

ARCHITECTURAL PRECAST CONCRETE

Form liners can be used to replicate stone textures
matching natural rock formations; fractured fins or
flutes; wood board markings; trapezoidal, wave, and
rib textures; sandblasted or bushhammered looks; and
stucco or masonry textures. The options with combination finishes, involving one or more basic finishing methods together with form liners, are almost infinite.
Ribbed or fluted panels demand considerable attention to detailing as panel sizes and distances between
openings must be a multiple of the rib spacing. Panel
joints should normally be in the bottom of a groove or
valley.
An important consideration is selecting the texture
and/or type of form liner best suited to the project. If
there are large wall expanses, a texture like fractured
fin with greater depth may give a more noticeable
appearance with deeper shadowing. Shallow flutes,
bushhammered, or subtle textures are often better for
relatively small areas. Concrete can be produced with
vertical ribs or striations in a range of sizes to suit a
particular structure and the distance from which it will
most often be seen. Overall, the cost of liners depends
on the ease of use and the number of reuses actually

SURFACE AESTHETICS

3.5.4 Form Liners and Lettering

3

Fig. 3.5.12
Washington County Fair Park
West Bend, Wisconsin;
Architect: BHS Architects;
Photo: BHS Architects.

obtained. Regardless of the form liner used, draft must
be incorporated to prevent chipping or spalling during
stripping of the unit from the mold.
The following rules should be observed when using
form liners:
• Limit depth of design to 1/2 to 1 in. (13 to 25 mm).
• In most cases, maintain a 1:8 draft on all indentation sides to prevent chipping and spalling during
stripping of the panel from the mold.
• Keep all edges and corners rounded or chamfered.
• Relief may be more than 1 in. (25 mm) if the depressed area is sufficiently wide.
Liner size and characteristics may require that an
architectural feature in the form of a demarcation
groove, recess, rib, or plain area be detailed to hide
joints between liners, or limit usage to within less than
the available width of the liner, or the liner joints should
be designed at form edges.
If the concrete is to be left smooth as-cast (that is,
without further treatment), its appearance will be
determined by the surface characteristics of the liner

material as well as by the chosen pattern or texture.
Variations in the absorbency of the form surface will
tend to produce corresponding variations in the color
of the concrete, a dark color being associated with water loss.
Sealed sandblasted wood, textured plywood, and
rough-sawn lumber are useful in creating rugged textures. (Resultant surface texture may also be obtained
by use of other liners reproducing this finish.) Roughsawn lumber is used for board-surface textured finishes where concrete color variations and rough edges are
acceptable. To provide the desired rural image, the insulated sandwich panels in Fig. 3.5.12 were cast using
custom form liners that created two different finishes.
The textures for the building’s upper portions were
molded from weathered barn boards, which produced
a close match to true wood. Panels on the lower portion resemble field stone, with the form liners molded
from limestone. The lower panels were stained on site
to give them a whitewashed look.
If preformed plastic form liners are selected, it is
good practice to describe the pattern and to include
a reference to the pattern and its manufacturer
specification.

ARCHITECTURAL PRECAST CONCRETE

|

161

3

SURFACE AESTHETICS
3.5.4 Form Liners and Lettering

Fig. 3.5.13
Philip Morris U.S.A., Cabarrus County, North Carolina;
Architect: MBA/Architects and Planners; Photo: Beckhard Richlan
& Associates.

(b)

(a)

Fig. 3.5.14(a) & (b)
Casa Club Bosque Real, Huixquilucan, Edo. de Mexico; Architect: Sordo Madaleno y Associados, S.C.;
Photos: Paul Citrón.

162

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.4 Form Liners and Lettering

(b)

3

(a)

Fig. 3.5.15(a) & (b)
Hilton Boston Logan Airport, East Boston, Massachusetts; Architect: Cambridge Seven Associates Inc.; Photos: ©2003Steve Rosenthal.

The 8 × 28 ft (2.4 × 8.5 m) insulated sandwich panels in Fig. 3.5.13 simulate the hand-hammered look
of fractured fins. The panels were produced by first
building a smooth-ribbed mold of wood and casting
one master panel from it with alternating directions
of diagonal fractured ribs. That panel was then hammered and sandblasted, and an elastomeric mold was
made of its hand-finished surface. This second mold
was used to cast the final panels. After demolding, the
rib surface was sandblasted to expose the aggregate to
the desired texture.
To replicate slate stone textures, rubber form liners
were reproduced from a natural slate quarry wall.
Different liner sections were rotated to avoid repetitive
patterns on the panels (Fig. 3.5.14[a]). Deep reveals
representing the joints between modular hewn stones
match the joints between panels. Interlocking lateral

ends avoided vertical joints (Fig. 3.5.14[b]).
The large panels for the hotel in Fig. 3.5.15 included
two window openings. The panels were cast in two
integral colors as well as two finishes. A brick-red
color is the building’s dominant color, accented by a
buff tone. A four-step process was used to create the
panels’ bushhammered ribbed pattern. First, a wood
mold was made detailing the ribs. Then a master mold
was made of concrete with an exposed-aggregate finish. Polyurethane was poured into the master mold to
obtain the form liner and then the colored concrete
was placed. Separated by reveals, ribbed sections were
combined with smooth areas, and the whole panel finally received an acid-etched finish. The panels were
alternated, with the diagonal direction of the ribs and
the colors varying from panel to panel. The project required 50 different form liner mats.

ARCHITECTURAL PRECAST CONCRETE

|

163

3

SURFACE AESTHETICS
3.5.4 Form Liners and Lettering

(a)

Fig. 3.5.16(a), (b) & (c)
Hearst Tower
Charlotte, North Carolina;
Architect: Smallwood, Reynolds, Stewart, Stewart & Associates Inc.

(c)

164

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.4 Form Liners and Lettering

The designer’s goal was to blend the façade of the 659ft-tall (200 m) office tower in Fig. 3.5.16 with its surroundings and complement existing buildings. To achieve
this blend, specially designed art deco patterns were cast
into panels at the fourth level to break up the apparent

3

mass of the larger footprint of the lower the tower elevations (Fig. 3.5.16[b]). Molds for these detailed pieces,
which contain a high degree of surface relief, took nearly
two weeks to fabricate (Fig. 3.5.16[c]).

Fig. 3.5.16(b)

ARCHITECTURAL PRECAST CONCRETE

|

165

3

SURFACE AESTHETICS
3.5.4 Form Liners and Lettering

(a)
(b)
Fig. 3.5.17(a) & (b) Arizona Biltmore Parking Structure, Phoenix, Arizona; Architect: Nelsen Architects Inc.; Photos: Rod Eaves.

166

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.4 Form Liners and Lettering

(a)

3

Fig. 3.5.19(a) & (b)
250 Park Avenue, Winter Park, Florida;
Architect: Baker Barrios Architects Inc.;
Design Consultant: Associated Consulting
International; Photos: Phil Eschbach.

The parking structure in Fig. 3.5.17(a)
was designed to fit seamlessly into an
architecturally important luxury resort
built in 1929. The original building is
comprised of unique and distinctive,
custom concrete masonry. Intricate form
liners were created from the original
hotel block. Following the precedent of
the hotel, detail and patterning is concentrated on the vertical elements (Fig.
3.5.17[b]). Spandrels have scored joints
matching the exact size and depth of the
original masonry joints. The integrally
colored concrete was given an acid-wash finish.
The architectural essence of the chevron (Fig. 3.5.18)
was designed to enhance the office building’s verticality
while establishing textural shades and shadows.

(b)

The renovation of a very contemporary 1960s building
into a traditional style that responds to the high-profile,
yet quaint, character of the neighborhood resulted in
the use of detailed architectural precast concrete panels
to establish the architectural vernacular (Fig. 3.5.19[a]).
The use of precast concrete spandrels with the owner’s
motif as a repetitious theme along with “old style” tavern blend thin-brick inlay in the precast concrete panels
to integrate with the historical district brought new life
to an old building (Fig. 3.5.19[b]). With an 85% occupancy rate during the entire construction timeline, great
pains were taken to ensure those businesses who remained in the building were not inconvenienced.
Sculptural designs have been produced using sections
of foamed polystyrene or polyurethane as form liners
or inserts. Abstract patterns and deeply revealed designs with undercut edges can be shaped easily in these
materials, however these liners are typically single-use
only. Computer-controlled, hot-wire cutting devices
have made custom work available at moderate prices.
Fig. 3.5.18
101 Hudson
Jersey City, New Jersey;
Architect: Brennan Beer Gorman (BBG-BBGM)/Architects;
Photo: BBG-BBGM.

Elastomeric liners are useful for finely detailed, textured, or
profiled surfaces with some undercuts (negative drafts) because they greatly facilitate stripping. If other materials were
used for such detail, the forms would be virtually impossible
to strip. Liner size and module should be coordinated with
panel joints, rustication strips, and blockout size.

ARCHITECTURAL PRECAST CONCRETE

|

167

3

SURFACE AESTHETICS
3.5.4 Form Liners and Lettering

Thought should be given to the selection of the letter profile or cross-section. Observing the principles of
shades and shadows and selecting a profile will give
sharp, smooth, and regular shadows. Two profiles for
recessed letters are shown and their merits analyzed
in Fig. 3.5.20. Raised letters are fragile and subject
to chipping at traffic levels and significantly increase
forming costs.

VI
EW

Lettering
The application of lettering in concrete is no different
than that of any other incised element. Appropriate
draft or taper for stripping must be established for all
lettering unless characters are flexible or destructible.

(a)

(b)

The pattern for the letters is reversed in the mold. Note
the letter “c” that is part of the word “school (need to
look at the mold from the right) (Fig. 3.5.21[a]). The
erected panels are shown in Fig. 3.5.21(b)
Fig. 3.5.20 Recessed lettering.

Close-up view of lettering on a precast concrete panel with
right angle shoulders

Fig. 3.5.21(a) & (b) Letters are reversed in the mold. Note the
letter “c” that is part of the word “school”.
Centralia High School, Centralia, Illinois; Architect: FGM Architects
Engineers, Inc.; Photos: (a) Jim Lewis, (b) Max Rogers.

(b)
Square shoulders of the V-recessed
letter makes a sharp shadow, but
the broken surface of the back
causes an uneven shadow making
the letter appear irregular

Recessed letters with right angle
shoulders and flat back stand out
clearly, because the shadow cast
by the outer angle against the flat
back is strong and regular

(a)

Fig. 3.5.22(a) & (b)
Music Man Square, Mason City, Iowa; Architect: Bergland + Cram;
Photos: Boxwood/Bergland + Cram/Boxwood.

168

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.4 Form Liners and Lettering / 3.5.5 Sand or Abrasive Blasting

The visibility of letters, is to some extent, determined
by the background and the style of the letters. The use
of contrasting precast concrete finishes or staining the
back of recessed letters in a color contrasting with the
surface wall will enhance the visibility of the letters. In
addition, design elements smaller than 1/300th of the
viewing distance are difficult to “read” and tend to get
visually lost.
The affordability and flexibility of the precast concrete panels made possible the unique detailing of
the cornice and base bands of the museum honoring
Meredith Wilson, composer of The Music Man. The
cornice displays the lyrics of “Seventy-Six Trombones”
(Fig. 3.5.22[a]) while the musical staff and notes at
the wall base parody the melody line of the song (Fig.
3.5.22[b]).

3.5.5 S and or Abrasive Blasting
Sand or abrasive blasting of surfaces can provide all
three degrees of exposure (Fig. 3.4.5, page 152). This
process is suitable for exposure of either large or small
aggregates.
The degree of uniformity obtainable in a sandblasted
finish is generally in direct proportion to the depth of
material removal. A light sandblasting may look acceptable on a small sample, but uniformity is more difficult to achieve at full scale, particularly if the units are
sculptured.
Uniformity of depth of exposure between panels and
within panels is essential for achieving an acceptable
finish, as in all other exposed-aggregate processes, and
is a function of the skill and experience of the operator.
Different shadings and, to some extent, color tone will
vary with the degree of aggregate exposure.
A light blast will emphasize visible defects, particularly bugholes, and reveal defects previously hidden by the surface skin of the concrete. A light blast
does minimize crazing by removing the cement skin
at the surface of the concrete. The lighter the blast,
the more critical the skill of the operator, particularly
if the units are sculptured. Small variances in concrete
strength at the time of blasting may further complicate
results. Sculptured units will have air voids on vertical
and sloped returns that may be accentuated by a light
blast. If such air holes are of reasonable size, 1/8 to 1/4
in. (3 to 6 mm), it is strongly recommended that they

3

be accepted as part of the texture, because filling and
sack-rubbing is expensive and will nearly always cause
color differences.
To improve uniformity, the cement and sand colors
should be chosen to blend with the slightly “bruised”
color of the sandblasted coarse aggregate, as the matrix color will dominate when a light sandblast finish
is desired. With a light sandblasting, only some of the
coarse aggregates near the surface will be exposed.
With a medium or deep exposure, contrasting matrix
and coarse aggregate colors should be avoided if uniformity of color is desired.
Blasting will cause some frosting of the face of the
coarse aggregate, and softer aggregates will show
this to a greater extent beyond a medium exposure.
Frosting of the aggregate surface is more noticeable
on dark-colored aggregates that have an initial glossy
surface texture. This will produce a muted or frosted
effect, which tends to lighten the color and subdue the
luster of the aggregate. For example, white concrete
tends to become whiter when blasted. Depth of sandblasting should also be adjusted to suit the aggregate
and abrasive hardness. Soft aggregates tend to erode
at the same rate as the mortar. There is a tendency to
round off edges of soft aggregates during sandblasting and soften sharp edges and corners.
Type and grading of abrasives determine the surface
texture and should remain the same throughout the entire project. Experienced precasters will select suitable
sandblasting techniques and media—the specification
should concentrate on the required appearance.
Although sandblasting is generally specified as an
overall treatment it may be used to develop textured
patterns by means of special templates. Portions of
a panel can be left unblasted by making a shield of
wood, rubber, or sheet metal to fit over the panel and
cover those areas. Masking may be adopted for geometric patterning, or the technique can be employed
by artists in producing murals in concrete.
The time when sandblasting should take place is determined by scheduling, economics, visual appearance
desired, and hardness of the aggregate. However, all
surfaces should be blasted at approximately the same
age or compressive strength for uniformity of appearance. The concrete mixture used and the matrix strength
at time of blasting will affect the final exposure, as will
the gradation and hardness of the abrasive.

ARCHITECTURAL PRECAST CONCRETE

|

169

3

SURFACE AESTHETICS
3.5.5 Sand or Abrasive Blasting

Fig. 3.5.23
Cape Coral City Hall
Cape Coral, Florida;
Architect: Spillis Candela/DMJM;
Photo: Larry Kline, Spillis Candela/DMJM.

The project design in Fig. 3.5.23 focused
on the texture of the precast concrete
as it changed from a light sandblast to
horizontal ribs. In Fig. 3.5.24, special attention was taken to design the 10-story
parking structure to harmonize with the
office building sheathed in polished and
flame-finished granite. The precast concrete was lightly sandblasted to closely
match the flame-finished granite of the
office tower base.

Fig. 3.5.24
Carillon Parking Deck
Charlotte, North Carolina;
Architect: Thompson, Ventulett, Stainback &
Associates;
Photo: Brian Gassel/TVS.

170

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.5 Sand or Abrasive Blasting

3

(a)
(b)

Fig. 3.5.25(a) & (b) CNL Center, Orlando, Florida; Architect: HKS Inc.; Photos: BenTanner.com.

A total of 1574 medium sandblasted precast concrete
panels were used to clad the 14-story office building
in Fig. 3.5.25(a). The concrete mixture comprised pea
gravel, white sand, and cement along with a pigment
to give a pink granite color (Fig. 3.5.25[b]).

The four-story operations center and five-level parking structure (Fig. 3.5.26) are clad with 6 in. (150 mm)
thick panels that have two finishes developed from the
same concrete mixture. The predominate finish is a deep
retarded finish with the second finish of medium sand-

ARCHITECTURAL PRECAST CONCRETE

|

171

3

SURFACE AESTHETICS
3.5.5 Sand or Abrasive Blasting

Fig. 3.5.26
AmSouth Bank Riverchase Operations Center; Hoover, Alabama; Architect: Smallwood, Reynolds, Stewart, Stewart & Associates.

172

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.5 Sand or Abrasive Blasting

(a)

3

blast banding.
The panels on the 36-story skyscraper in Fig. 3.5.27(a)
are characterized by a striking blend of white cement,
light and heavy sandblast finishes, and four different
aggregates. The result is the appearance of natural
granite inlays in a field of traditional precast concrete
panels (Fig. 3.5.27[b]). A design feature was the incorporation of fire/smoke ribs directly into the precast
concrete column covers during the fabrication stage.
Fig. 3.5.27(a) & (b)
State Street Financial Center
Boston, Massachusetts;
Architect: TRO Jung/Brannen Associates Inc.;
Photos: Peter Vanderwarker.
(b)

ARCHITECTURAL PRECAST CONCRETE

|

173

3

SURFACE AESTHETICS

3.5.5 Sand or Abrasive Blasting / 3.5.6 Acid Etching

A sandblasted finish is not widely used to achieve a
deep, heavy texture because of the time and labor associated with deep exposure. Unless it is the intent of the
architect to achieve a severely weathered look, deep
exposed-aggregate finishes are more readily achieved
with other methods. For example, to obtain a medium
or deep exposure with a sandblasted appearance, retarders may be used initially followed by sandblasting
to obtain a matte finish. This approach reduces blasting time and lessens the abrasion of softer aggregates.
Using sandblasting to achieve the final texture allows
for correction of any variations in exposure, so this
method can result in a more uniform surface. The end
result is a matte finish, as opposed to a brighter finish
achieved with water blasting.

a light sandblast texturing. Where aggregates are to
be exposed to a considerable depth, only acid-resistant siliceous aggregates, such as quartz and granite,
should be used. Carbonate aggregates, such as limestone, dolomite, and marble, may discolor or dissolve
due to their high calcium content. The aggregates on
an acid-etched surface present a clean or bright look.
However, after normal weathering, the aggregates lose
this brightness and will closely resemble their original
condition.
All surfaces should be acid-etched at approximately
the same age or compressive strength for uniformity
of appearance.
Acid etching of concrete surfaces will result in a fine,
sandy texture with retention of detail. When the acid
etching is light or used on a large, plain surface, concentrations of cement paste, under and over etching
of different parts of a concrete surface and variation
in sand color or content may cause some uniformity
problems.

Exposed aggregates can be brightened by washing
with a mild acid solution, which removes the dull cement film remaining from some exposure techniques,
such as sandblasting and retardation.

3.5.6 A
 cid Etching

There is a minimum depth of etch that is required to
obtain a uniform surface. Attempts to go any lighter
than this will result in a blotchy panel finish. This depth
will expose sand and only the very tip of the coarse aggregate. It is difficult to achieve a totally uniform light
exposure on a highly sculptured panel. This is due to
the acid spray being deflected to other areas of the

Acid etching is most commonly used for light to medium exposures. Acid etching dissolves the surface cement paste to reveal the sand with only a very small
percentage of coarse aggregate being visible. An
acid-etched finish is typically used to produce a fine
sand texture closely resembling natural stones such
as limestone or sandstone. It is often substituted for

Fig. 3.5.28(a) & (b)
Walsh Library at Seton Hall University
South Orange, New Jersey (1994);
Architect: Skidmore, Owings & Merrill;
Photos: Eduard Hueber/Archphoto.com.

(a)

174

|

ARCHITECTURAL PRECAST CONCRETE

(b)

SURFACE AESTHETICS

3.5.6 Acid Etching

3

achieved by banding at every 7 ft (2.1 m) datum. In
any case, a light acid-etched finish is not going to be
as uniform as an exposed-aggregate finish. Also, an
acid-etched finish is more difficult to patch than many
of the deeper texture finishes.
Vertical rustications or reveals should extend to the
bottom of the unit to avoid the potential of a deeper
etch on the bottom flat band of a unit. If not extended,
acid collects in the reveal and can run down and streak
the bottom band.
The all–precast concrete structure features modules,
measuring 4 stories high and 10 ft (3 m) wide (Fig.
3.5.30). This module allowed the façade to curve in response to the shape of the nearby Renaissance Center.
A mixture of red granite aggregate, red sand, and
red-pigmented cement was used to blend the panels’
coloration with neighboring brick and bronze-glass
buildings. The panels were given a medium acid etch
to create a smooth finish and expose the richly colored
texture.
Fig. 3.5.29
Northwestern University McCormick Tribune Foundation Center,
Evanston, Illinois; Architect: Einhorn Yaffee Prescott; and
Griskelis Young Harrell, Associate Architect; Photo: Nick Merrick
©Hedrich Blessing.

panel, particularly at inside corners. This may, however,
be acceptable if the sculpturing creates differential
shadowing.
Figure 3.4.5 (page 152) shows a light, medium, and
deep acid-etched finish with the same concrete mixture used for the retarded and sandblasted finishes.
With light textures, the color compatibility of the cement and the aggregates become more important to
avoid a mottled effect. The complexion of the precast
concrete used in Fig. 3.5.28(a) responds to the color
and texture of the surrounding academic buildings.
The color was selected to simulate the color and texture of Indiana limestone and the panels were finished with a light acid-etch. The interior column covers for the rotunda were also given a light acid-etch
(Fig. 3.5.28[b]). The selection of architectural precast
concrete with granite insets on the exterior allowed a
great range in the expression of architectural details,
and wider possibilities for functional articulation. The
broadcast media center in Fig. 3.5.29 is sheathed with
lightly acid-etched panels carefully modulated to pick
up the scale of smaller, neighboring buildings. This was

Fig. 3.5.30
Jefferson Avenue Parking Structure, Detroit, Michigan;
Architect: Neumann/Smith & Associates;
Photo: Hedrich Blessing.

ARCHITECTURAL PRECAST CONCRETE

|

175

3

SURFACE AESTHETICS
3.5.6 Acid Etching

(a)

(b)

(c)

Fig. 3.5.31(a), (b) & (c)
The Minneapolis Convention Center, Minneapolis, Minnesota;
Architect: Setter, Leach & Lindstrom Inc.; The Leonard Parker
& Associates; and Loschky, Marquardt & Nesholm
Photos: Heinrich Photography.

176

|

ARCHITECTURAL PRECAST CONCRETE

The two-story convention center (Fig. 3.5.31[a]), provides over 350,000 ft2 (32,500 m2) of exhibit space
in three domed exhibition halls and a multipurpose
ballroom. The exterior precast concrete is a sandstone
red color with a deep acid-etched finish along with a
polished concrete band at the base of the wall. Bluegreen tiles are inlaid in squares and circles on certain
key panels for color and texture contrast (Fig. 3.5.31[b]
and [c]). The insulated sandwich panels are horizontally
inscribed with a series of reveals to suggest the heavier
jointing and rustication of traditional stone buildings.

SURFACE AESTHETICS

3.5.7 Multiple Mixtures and Textures within a Single Unit

3.5.7 M
 ultiple Mixtures and Textures
within a Single Unit
Design flexibility is possible in both color and texture
of precast concrete by manipulating aggregates and
matrix colors, size of aggregates, form liners, finishing
processes, and depth of exposure in the same unit. This
textural flexibility allows designers to use combinations
of different finishes using the same or different concrete mixtures, within a single precast concrete unit.
Multiple-finishing techniques offer an economical, yet
effective, way to heighten aesthetic interest through
the use of tones and texture in façade treatments. The
use of combination finishes means the designer must
make an early decision to ensure that the overall concept allows for the change in finish color and texture. A
suitable rustication (that is, some demarcation) needs
to be detailed to separate the different colors and/or
finishes. The importance of the separation depends on
the specific types of finishes involved.

3

Fig. 3.5.32 Multiple mixtures
and textures.

Multiple mixes with sandblasted
and acid-etched finishes.

Light and deep sandblast with
granite and terra-cotta inserts.

Multiple mixes with acid-etched
finish.

Samples should be used to assess the transition between adjacent finishes. Bushhammering and, to a
lesser degree, sandblasting can be stopped fairly easily
along specific lines and may not require the need for the
demarcation features as described in Section 3.3.3.
There are two approaches for using multiple mixtures
(two or more different facing mixtures in the same
panel). With one approach, the first mixture is placed
within an area bounded by a raised demarcation strip
that equals the thickness of the face mixture. Before
initial set of the concrete, the mold surface around the
first cast is carefully cleaned, and the second mixture
is placed and vibrated. It is important that the second
mixture be placed and the concrete consolidated prior
to initial set of the first concrete mixture.
Another approach features a two-stage or sequential
casting procedure discussed in Section 3.3.9, which incurs added cost. In this option, one part of the panel,
such as a medallion, is cast first from one mixture and,
after curing, is set into the full mold and cast into the
total panel using a second mixture. This method was
used to cast precast concrete panels cost effectively
in three finishes (Fig. 3.3.37, page 134). The retarded
rosebud quartzite panel section was cast separately,
set in a mold, and then the remainder of the concrete
was cast around it.
Examples of projects that effectively use multiple mixtures and finishes are shown in Fig. 3.5.32.

Multiple mixes with retarded
and sandblasted finishes.

Natural stone and sandblasted
finish.

Thin brick and acid-etched finish.

Form liner and sandblasted finish.

Multiple mixes with acid-etched
and form liner finishes.

Thin brick, acid-etching and form
liner.

ARCHITECTURAL PRECAST CONCRETE

|

177

3

SURFACE AESTHETICS

3.5.7 Multiple Mixtures and Textures within a Single Unit

Fig. 3.5.33
Westings Corporate Center I, Naperville, Illinois; Architect: Opus Architects & Engineers, Inc.

Multiple finishes provide a variety of textures at eye
level, not only to add interest upon approach of the
building but also to visually ground the building with
its darker mass from a distance (Fig. 3.5.33). This was
accomplished by using a medium gray matrix that has

178

|

ARCHITECTURAL PRECAST CONCRETE

been lightly sandblasted to contrast with the dark charcoal aggregate that has been exposed using a retarder
in horizontal bands. A third texture was achieved by
using a vertically revealed form liner in the medium
gray matrix panels.

SURFACE AESTHETICS

3.5.7 Multiple Mixtures and Textures within a Single Unit

3

Fig. 3.5.34
Monarch Place
Boston, Massachusetts;
Architect: TRO Jung Brannen Associates, Inc.;
Photo: ©1987 Steve Rosenthal.

An acid-etched finish cannot easily be applied to only
a portion of a unit with alternating surfaces of retarded
and acid-etched precast concrete, a reveal or raised
demarcation feature is necessary to keep the retarder
from spreading to the area to be etched [Fig. 3.3.6(a)]
(page 114).

ture and the other a darker, pebbled texture produced by
retarding and water-washing the red granite aggregate.
A series of ridges and reveals creates the patterned effect
of these surfaces (Fig. 3.5.34). The corner pieces were
cast as complete units to provide a smooth, clean corner
without a break, which kept the rustications aligned.

The same mixture proportions were used to develop two
textures: one, a smooth light acid-etched pale pink tex-

To emulate historical French limestone construction
of turn-of-the-century buildings, multiple joint lines

ARCHITECTURAL PRECAST CONCRETE

|

179

3

SURFACE AESTHETICS

3.5.7 Multiple Mixtures and Textures within a Single Unit

Fig. 3.5.35
55 Farmington, Hartford, Connecticut;
Architect: Thompson, Ventulett, Stainback & Associates, (TVS);
Photo: Brian Gassel/TVS.

180

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.7 Multiple Mixtures and Textures within a Single Unit

were incorporated in the acid-etched portions (Fig.
3.5.35). The spandrels, however, have a retarded finish that–through its differing texture–accentuates the
horizontal aspects of the building.
The luxury 27-story condominium features architectural precast concrete panels with local limestone
aggregates with a light sandblast finish and two face
mixtures (Fig. 3.5.36). The project’s two colors include
(a)

(b)

Fig. 3.5.36(a) & (b)
Villa D’Este Condominiums
Houston, Texas;
Architect: Ziegler Cooper Architects;
Photos: Aker/Zvonkovic Photography.

3

a soft, warm wheat color for the facing with a light,
white/vanilla color for the “shoulders” and trim on the
edges. Adding to this was horizontal fluting and textures created with form liners to produce 2 ft (0.6 m)
bands of rougher finishes. Adding color, texture, and
pattern reduced the massiveness of the building.
Several textures were cast into the precast concrete to
add shadow and life to the façade, including rock face,
stippled, smooth, and striated finishes (Fig. 3.5.37).
Visual interest and a unified structure can be obtained by composing harmonious patterns
into themes using a palette consisting of the
form and rustication lines of the surface, and
the texture and color of the precast concrete.
Precast concrete has extensive capabilities to
create textural options. Different shapes, a
variety of aggregates, and precast concrete’s
ability to form a variety of textures offer endless architectural variations.

3.5.37
Thomas F. Eagleton United States Courthouse;
St. Louis, Missouri
Architect: Hellmuth, Obata & Kassabaum, P.C.;
Photo: Timothy Hursley.

ARCHITECTURAL PRECAST CONCRETE

|

181

3

SURFACE AESTHETICS
3.5.8 Tooling or Bushhammering

3.5.8 T ooling or Bushhammering
Concrete can be mechanically spalled or chipped
with a variety of hand and power tools to produce
an abraded, exposed-aggregate texture. Each type of
tool produces a distinctive surface effect and a unique
shade of concrete color. All tooling removes a layer of
hardened concrete while fracturing larger aggregates
at the surface. It produces an appearance somewhat
different from other types of aggregate exposure. The
color of the aggregate, but not necessarily the aggregate shape, is revealed. This finishing technique is most
suitable for flat or convex surfaces, and is more labor
intensive than most other finishing processes. All surfaces should be tooled at approximately the same age
or compressive strength for uniformity of appearance.
Pneumatic or electric tools may be fitted with a
bushhammer, a flat or dentated chisel with one to six
teeth, a crandall, or multiple pointed attachments (Fig.
3.5.38[a]. The type of tool will be determined by the
desired surface effect. The finish obtained can vary
from light scaling to deep bold texture (Fig. 3.5.39[a],
[b], and [c]). Bushhammered finishes affect the appearance, color, and brightness of the aggregate. Color
tends to be lightened by the fracturing, which on dark
materials has a dulling effect, but it often improves the
light grayish and, in particular, white tones. By increasing or decreasing the shadow content of the texture,
tooling alters the panel reflectancy and changes the
tone value. Scaling produces a fine ribbed effect, rather than a deeply chipped texture. The bushhammer
produces a rougher texture, fracturing aggregate, and
removing up to 3/16 in. (5 mm) of material. A chiseled or
pointed tool fractures and accentuates the coarse aggregate and may remove as much as 3/4 in. (19 mm) of
concrete surface. Chisel-type tools are better for fracturing across aggregate particles, while pointed tools
Fig. 3.5.38

(a)

182

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 3.5.39(a), (b) & (c) Depths of exposure.
(a)

Light

(b)

Medium

(c)

Deep

tend to dig into the matrix. Bushhammering or tooling
should follow a specific direction to obtain a consistent
texture.
Although a dense, fully-graded concrete mixture is
desirable, bushhammering has been successfully applied to gap-graded concrete. Natural gravels are inclined to shatter, leading to bond failure and loss of
aggregate particles; it is preferable to use a crushed
aggregate concrete. Aggregates such as quartz and
granite are difficult to bushhammer uniformly because
of their hardness, and they may fracture into, rather
than across, the concrete surface. Aggregates such as
dolomite, marble, calcite, and limestone are softer and

SURFACE AESTHETICS

3.5.8 Tooling or Bushhammering

3

(a)

Fig. 3.5.40 (a) & (b)
Bristol Myers Squibb Corporation
Mexico, D.F., Mexico;
Architect: Migdal Arquitectos;
Photos: (a) Alberto Moreno, (b) Paul Citrón.

(b)

more suitable for bushhammered surfaces. The comb
chisel is suitable only for use with softer aggregates.
Concrete containing soft aggregates cannot be satisfactorily point-tooled.
Bushhammering at outside corners may cause jagged edges. If sharp corners are desired, bushhammering should be held back from the corner a distance of
1 to 2 in. (25 to 50 mm) or more. It is quite feasible
to execute tooling along specific lines. If areas near
corners are to be tooled, this usually is done by hand
because tools will not reach into inside corners, making this operation more expensive. Chamfered corners
are preferred with tooled surfaces and a 1 in. (25 mm)
chamfer may be tooled with care.
The façade elements in Fig. 3.5.40(a) and (b) were
finished imitating a textured stone and were done
with hand pneumatic chisel hammers. The textured
finish surfaces resulted in excellent volume contrasts,
where the handmade chiseling gave each panel its individuality without loosing the overall uniformity. The
face mixture used combined orange- brown ochre and
white natural stone aggregates with white cement.

ARCHITECTURAL PRECAST CONCRETE

|

183

3

SURFACE AESTHETICS
3.5.8 Tooling or Bushhammering

Figure 3.5.41 shows a close-up of a bushhammered
rib panel made with a yellow marble aggregate and
white cement. The panels in Fig. 3.5.42(a) and (b) feature a bold texture of bushhammered diagonal ribs
that alternate in direction at each horizontal band. The
texture reveals much of the green aggregate and provides striking shade and shadow effects on the curved
surfaces. The parapet panels also feature inset granite
accents and narrow notches suggesting crenulations.

Specifications for uniformity or non-uniformity of
tooled finishes are extremely difficult to write and assistance should be sought from the precaster who is
providing the tooled finish being specified.
Tooling removes a certain thickness of material, 3/16
to 3/8 in. (5 to 10 mm) on an average from the surface
of the concrete, and may fracture particles of aggregate causing moisture to penetrate the depth of the
aggregate particle. For this reason the minimum cover
to the reinforcement should be somewhat larger than
normally required. It is sometimes recommended that
2 in. (50 mm) of cover be provided (prior to tooling).
(b)

Fig. 3.5.41

(a)

Fig. 3.5.42(a) & (b)
Michelin North American
Corporate Headquarters
Greenville, South Carolina;
Architect: Odell Associates, Inc.;
Photos: Odell Associates, Inc.

184

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.8 Tooling or Bushhammering

Fig. 3.5.43

Fig. 3.5.44 Atlanta Central Library
Atlanta, Georgia; Photo: Image courtesy of Marcel
Breuer papers, 1920-1986 in the Archives of
American Art, Smithsonian Institution.

HAMMERED RIB OR FRACTURED FIN: A hammered
rib or fractured fin finish may be produced by casting
ribs on the surface of the panels and then using a hammer or bushhammer tool to break the ribs and expose
the aggregate. The profile selection should be considered in relation to: (1) readability at varying distances,
(2) the area of wall on which the finish is to be used in
order to achieve the desired effect, and (3) the ability
of the selected aggregate to penetrate the rib. Figure
3.5.43 shows a close-up of a fractured fin panel made
with a yellow marble aggregate and white cement.
The effect is a bold, deeply textured surface. Rib size
measured at the outer face should be a maximum of 1
in. (25 mm), as larger sections are difficult to fracture.
The ribs should not be narrower than 5/8 in. (16 mm)
or they may break off at their base without leaving any
of the rib projecting. The ribs may be hammered from
alternate sides, in bands, to obtain uniformity of cleavage, or randomly, depending on the effect required.
The hammering technique employed, whether carried
out in bands or in a random pattern, will alter the final
appearance and should be specified by the designer—
it may require a number of samples. There should be
a definite plan, even with so-called random pattern
because, unless care is exercised, an uneven shading
effect on the concrete surface may be produced.

3

Fig. 3.5.45

The diagonal striated pattern of the panels in Fig.
3.5.44 were designed for maximum color and texture.
Light-gray limestone and white river gravel were used
for color control. A warm buff sand added color to
the gray matrix. Before finishing, the ribs were 3/8 in.
(10 mm) tall and 3/8 in. (10 mm) wide at the base and
repeated at 1/2 in. (13 mm) intervals. The striated ribs
were bushhammered to expose the aggregates. The
size of the ribs and the size of the aggregate were
carefully coordinated to achieve maximum coarse aggregate exposure. Figure 3.5.45 shows a close-up of
the fractured fin.
Because this finish is labor intensive, it is expensive,
but may be justified if the panels will be viewed up
close on ground level walls or interior walls. On upper
floors a similar effect may be achieved at much less
cost by retarding or sandblasting the ribs. An effect
similar to a fractured rib finish may be achieved less
expensively by using forms or form liners that simulate
the fractured ribs. The exposed, weathered look can
then be achieved by chemical retardation or by sandblasting the rib surface (see Section 3.5.4). The panels
can have a flat border area to accommodate variations
in panel sizes, thus eliminating the need for any bulkheading in the ribbed area.

ARCHITECTURAL PRECAST CONCRETE

|

185

3

SURFACE AESTHETICS
3.5.9 Sand Embedment

3.5.9 S and Embedment
When bold and massive architectural qualities are desired, cobble stones ranging from 11/2 to 8 in. (38 to
200 mm) in diameter (Fig. 3.5.46), or large, thin slices
of stone such as fieldstones or flagstones (Fig. 3.5.47
[a] and [b]), may be exposed by the sand embedment
technique. These large stones must be hand placed in
a sand bed, or other special bedding material, at the
bottom of the mold to a depth that keeps the backup

Fig. 3.5.46

Fig. 3.5.47(a) & (b)
Bryan University Center
Duke University
Durham, North Carolina;
Architect: Hayes, Howell and Associates;
Photos: Hayes-Howell, PA.

(a)

186

|

ARCHITECTURAL PRECAST CONCRETE

concrete 25 to 35% of the stone’s diameter from the
face. This technique reveals the facing material and produces the appearance of a mortar joint on the finished
panel. Spandrel units in Fig. 3.5.48(a) are composed of
gray cement, local aggregates, and red/brown sand.
Natural red/orange-colored sandstone is cast into areas
of the panels, providing a pronounced texture and uncommon visual effect. Slabs of sandstone were broken
and the pieces were hand-placed, and then sand was
sifted into the joints (Fig. 3.5.48[b]).
Stones should be dense and evenly distributed on all
surfaces. This is particularly important around corners,
edges, and openings, as well as on the flat surfaces.
To help achieve uniform distribution and exposure, all
stones should be approximately one size. When facing
materials are of mixed colors, their placement in the
mold must be carefully checked for the formation of
unintended patterns or local high incidence of a particular color. When using some stones, if it is the intention to expose a particular facet of the stone, placing
should be checked with this in mind before the backup

(b)

SURFACE AESTHETICS

3.5.9 Sand Embedment / 3.5.10 Clay Product-Faced Precast Concrete

concrete is placed. Care should
be taken to ensure adequate
cover to reinforcement following
exposure.
Sample panels are essential for
this technique, if only to ensure
a compaction method that provides for full compaction without
dislodgement of the stones. It is
desirable that panels having this
embedment technique and using
large size stones be provided with
a margin, generally smooth, offform around all edges of the panel face. If that is not provided and
the aggregates are taken right up
to the arris of the panel, then subsequent washing and brushing
may give a non-uniform, torn appearance to the arris.

Fig. 3.5.48(a) & (b)
H&R Block Services Center
Kansas City, Missouri;
Architect: Berkebile Nelson Immenschub McDowell Architects;
Photos: ©2000 BNIM Architects.

3

(a)

3.5.10 Clay Product-Faced
Precast Concrete
3.5.10.1 General

(b)

Clay product-faced precast concrete is being used
increasingly today as another choice to obtain an aesthetic façade while blending in with surrounding structures. It gives the architect the flexibility to combine the
pleasing visual appearance of traditional clay products
with the strength, versatility, and economy of precast
concrete. Among the types of materials that can be
embedded in the precast concrete are brick, ceramic
tile, porcelain, and architectural terra cotta. These clay
product-facings may cover the exposed panel surface
entirely or only part of the concrete face, creating accents. The use of clay product–faced precast concrete

ARCHITECTURAL PRECAST CONCRETE

|

187

3

SURFACE AESTHETICS

3.5.10.1 General / 3.5.10.2 General considerations

panels began in the early- to middle-1960s. These early projects have not needed tuckpointing or sealing.
The combination of precast concrete and clay products offers several important benefits over site-laid-up
masonry.
Precasting techniques allow complex and intricate
details such as arches, radii, ornate corbels, and numerous bonding patterns to be incorporated into the
finished panel (Fig. 3.5.49). This freedom of aesthetic
expression could not economically be accommodated
with site-laid-up masonry. A prefabricating approach
ensures that high-priced and time-consuming building skills are transferred to the controlled conditions
of the plant and away from the critical path of on-site
activities.
Precasting also allows a high level of dimensional
precision and quality control. Concrete mixtures and
batching, together with curing conditions, can be
tightly controlled, whereas site-laid masonry may have
variable curing and mortar qualities.
Plant production provides for year-round work under
controlled temperature conditions, negating any on-site
delays due to inclement weather or incurring the expense
of on-site weather protection. It also allows the structure to be winterized in advance, with floor topping and
finishing trades continuing without any weather delays.
Clay product-faced precast concrete can eliminate the
need for costly on-site scaffolding and greatly reduce
the duration of masonry cladding time. Also, site disturbance, construction debris, and use of toxic cleaners are
reduced. Precast concrete design allows gravity loads to
be located at columns, eliminating expensive lintels and
mid-span loading on structure. Brick-faced precast concrete panels eliminate dovetail anchors, flashing, weep

Fig. 3.5.49

188

|

ARCHITECTURAL PRECAST CONCRETE

holes, and the need for lintels.
Panel configurations include a multitude of shapes
and sizes: flat panels, C-shaped spandrels, soffits, arches, and U-shaped column covers. Repetitive use of any
particular shape also lowers costs dramatically. Returns
on spandrels or column covers may be produced by the
sequential (two-stage) casting method or as a single
cast depending on the height of the return. Panels may
serve as cladding or may be loadbearing, supporting
floor and roof loads.

3.5.10.2 General considerations
Structural design, fabrication, handling, and erection
considerations for clay product–faced precast concrete
units are similar to those for other precast concrete
walls panels, except that consideration must be given
to the dimensional layout of the clay product material
and its embedment in the concrete. The physical properties of the clay products must be compared with the
properties of the concrete backup. These properties include the coefficient of thermal expansion, modulus of
elasticity, and volume change due to moisture, along
with strict adherence to tight dimensional tolerances.
For design purposes, clay product–faced precast concrete panels may be designed as concrete members
that neglect the structural action of the face veneer.
The thickness of the panel is reduced by the thickness
of the veneer, and design assumptions exclude consideration of differential shrinkage or differential thermal
expansion. However, if the panel is to be prestressed,
the effect of composite behavior and the resulting
prestress eccentricity should be considered in design.
Reinforcement of the precast concrete backup should
follow recommendations for precast concrete wall
panels relative to design, cover, and placement.
The height and length of the panels should be multiples of nominal individual masonry unit heights and
lengths for effective cost control in the precast concrete
production process. The actual specified dimensions
may be less than the required nominal dimensions by
the thickness of one mortar joint. For economical production, the precaster should be able to use uniform
and even coursing without cutting any units vertically
or horizontally except as necessary for precast panel
joints and bond patterns. The PCI Standard for embedded brick in precast concrete panels should be specified to ensure size uniformity, long term durability and
material compatibility.

SURFACE AESTHETICS

3.5.10.2 General considerations

PCI Standard for Thin Brick
The objective of this standard is to outline material standards
and specification criteria for brick manufacturers to meet
when supplying materials to precast concrete manufacturers.
The intent is to establish acceptable dimensional tolerances
and consistent testing standards for brick embedded in precast concrete systems. The brick manufacturers must confirm
through the provision of independent test results that their
brick products comply with the PCI Standard. The PCI Standard
should appear in all specifications as the new, approved industry standard. Brick manufacturers have agreed to promote the
compliance of their brick with this new standard.
The following parameters have been established based on
the successful use of embedded brick in precast concrete projects. The parameters set forth for use in these proposed standards are attainable brick properties that have been derived
with input from brick manufacturers, precasters, engineers,
and architects, as well as consideration of existing test results.
A. Thin Brick Units: PCI Standard, not less than 1/2 in. (13 mm)
nor more than 1 in. (25 mm) thick with an overall tolerance of
plus 0 in., minus 1/16 in. (+0 mm, -1.6 mm) for any unit dimension 8 in. (200 mm) or less and an overall tolerance of plus 0 in.,
minus 3/32 in. (+0 mm, -2.4 mm) for any unit dimension greater than 8 in. (200 mm) measured according to ASTM C 67.
1. F ace Size: Modular, 21/4 in. (57 mm) high by 75/8 in.
(190 mm) long.
2. F ace Size: Norman, 2 1/4 in. (57 mm) high by 115/8 in.
(290 mm) long.
3. Face Size: Closure Modular, 35/8 in. (90 mm) high by 75/8
in. (190 mm) long.
4. F ace Size: Utility, 35/8 in. (90 mm) high by 115/8 in. (290
mm) long.
5. F ace Size, Color, and Texture: [Match Architect’s
approved samples] [Match existing adjacent
brickwork].
a. 
6. S pecial Shapes: Include corners, edge corners, and end
edge corners.
7. C
 old Water Absorption at 24 hours: Maximum 6%
when tested per ASTM C 67.
8. E fflorescence: Provide brick that has been tested according to ASTM C 67 and rated “not effloresced.”
9. O
 ut of Square: Plus or minus 1/16 in. (+/- 1.6 mm)
measured according to ASTM C 67.
10. Warpage: Consistent plane of plus 0 in., minus 1/16 in.
(+0, -1.6 mm).

3

11. Variation of Shape from Specified Angle: Plus or minus
1 degree.
12. Tensile Bond Strength: Not less than 150 psi (1.0 MPa)
when tested per modified ASTM E 488. Epoxy steel
plate with welded rod on a single brick face for each
test.
13. F reezing and Thawing Resistance: No detectable
deterioration (spalling, cracking, or chafing) when
tested in accordance with ASTM C 666 Method B.
14. Modulus of Rupture: Not less than 250 psi (1.7 MPa)
when tested in accordance with ASTM C 67.
15. Chemical Resistance: Provide brick that has been tested
according to ASTM C 650 and rated “not affected”.
16. Surface Coloring: Brick with surface coloring shall withstand 50 cycles of freezing and thawing per ASTM C
67 with no observable difference in applied finish when
viewed from 20 ft (6 m).
17. Back Surface Texture: [Scored], [Combed], [Wire
roughened], [Ribbed], [Keybacked], [Dovetailed].
Test sample size and configuration shall conform to the following parameters in order to validate compliance by brick
manufacturer with PCI Standard for use in embedded brick
precast concrete systems:
1. Minimum number of tests specimens: Comply with appropriate specifications except for freeze-thaw and tensile
bond strength tests on assembled systems.
2. Minimum number of test specimens for freeze-thaw and
tensile bond strength test: Ten (10) assembled systems
measuring 8 x 16 in. (200 mm x 405 mm) long with the
brick embedded into the concrete substrate (assembled
system). The ten (10) assembled systems are divided into
5 Sample A assemblies and 5 Sample B assemblies. The
precast concrete substrate shall have a minimum thickness
of 21/2 in. (63 mm) plus the embedded brick thickness.
The precast concrete shall have a minimum compresssive
strength of 5000 psi (34.5 MPa) and 4 to 6% entrained air.
The embedded brick coursing pattern for testing purposes
shall be modular size brick on a half running bond pattern
with a formed raked joint geometry of no less than 3/8 in. (9
mm) wide and a depth no greater than 1/4 in. (6 mm) from
the exterior face of the brick. One brick from the center of
each Sample A assembly shall be tested for tensile bond
strength, Item #12. Each Sample B assembly shall first be
tested for freeze-thaw resistance, Item #13 and then one
brick from the center of each Sample B assembly shall be
tested for tensile bond strength, Item #12.

ARCHITECTURAL PRECAST CONCRETE

|

189

3

SURFACE AESTHETICS

3.5.10.2 General considerations / 3.5.10.4 Clay product selection

The appearance of clay product–faced precast concrete panels is achieved principally by the selected clay
product, with type, size, and texture contributing to
overall color. Also, the degree to which the clay product units are emphasized will depend upon the profile
and color of the joint between units. The Brick Institute
of America (BIA) recommends concave joints in all masonry projects. Due to forming requirements and material tolerances, it is preferable that joints between
clay products be not less than 3/8 in. (10 mm).
The joints between panels are usually butt joints.
Corners are usually achieved by using brick returns equal
to the length of the brick module. The final element in
the appearance of the panel is the 5000 psi (34.5 MPa)
concrete used in the joints. Hand-tooled joints may be
simulated by form liners or joints may be tuckpointed
after forms are stripped, however this may add to the
cost and maintenance of the panel.
The contract documents should clearly define the
scope of clay product sizes, coursing patterns, and
placement locations. Both stack and running bond
patterns have been used widely in precast concrete
panels. These patterns can be interchanged with soldier courses, basket weave, or herringbone patterns.
Running bond patterns are typically less costly and visually more appealing when courses start and finish
with half or full brick. This approach avoids cutting and
allows matching adjacent spandrels or column covers.
Also, providing a narrow strip of exposed concrete at
the edges of the panel helps reduce the visual impact
and potential difficulty in aligning brick joints between
precast concrete units. Vertical alignment of joints, especially with stack bond, requires close clay product
tolerances or cutting of brick to the same length.

3.5.10.3 Clay product properties
Physical properties of clay products vary depending
on the source of clay, method of forming, and extent of firing. Table 3.5.1 shows the range of physical
properties of clay products. Because clay products are
subject to local variation, the designer needs to obtain
information on the specific brick being considered to
ascertain if the variations are acceptable.
As the temperature or length of the burning period is
increased, clays burn to darker colors, and compressive
strength and modulus of elasticity are increased. In general, the modulus of elasticity of brick increases with compressive strength to a compressive value of approximately
5000 psi (34.5 MPa); after that, there is little change.

3.5.10.4 Clay product selection
Precasters should be consulted early in the design
stage to determine available colors, textures, shapes,
sizes, and size deviations of clay products, as well as
manufacturing capability for special shapes, sizes, and
tolerances. The specification should identify the color,
size, and manufacturer of the clay product. Usually the
precast concrete producer buys the clay products and
knows which products are able to conform to the PCI
Standard for embedded brick in precast concrete.
PCI Standard thin-brick veneer units 1/2 to 1 in. (13 to
25 mm) thick are typically used and are available in various sizes, colors, and textures. Thin brick conforming to
PCI Standard are actually a tile and have lower water
absorption than conventional brick. In addition, thin
brick is less susceptible than conventional brick to freezing and thawing issues, spalling, and efflorescence.
Stretcher, corner, or three-sided corner units are typi-

Table 3.5.1. Range of physical properties of clay products.

Type of Unit

Compressive
Strength, psi

Modulus of
Elasticity, psi

Brick

3000 – 15,000

1.4 – 5.0 × 106

Ceramic or Quarry tile 10,000 – 30,000
Glazed wall tile

8000 – 22,000

1.4 –5.0 × 106

Terra cotta

8000 – 11,000

2.8 – 6.1 × 106

Note: 1 psi = 0.006895 MPa.

190

|

7 × 106

ARCHITECTURAL PRECAST CONCRETE

Tensile
Strength, psi

Coefficient of
Thermal
Expansion
in./in. °F
4 × 10-6

Approx. 0.1
compressive
strength

2.2 – 4.1 × 10-6
4.0 – 4.7 × 106
4 × 10-6

SURFACE AESTHETICS

3.5.10.4 Clay product selection

3

Fig. 3.5.50 Thin brick units.
(a)

Stretcher

(b)

Corner

cally available in a variety of color ranges (Fig. 3.5.50).
The face sizes normally are the same as conventional
brick and, therefore, when in place, yield the aesthetics
of a conventional brick masonry wall with the superior
performance of precast concrete.
The most common brick face size is the modular. The
utility face size is popular for use in large buildings because productivity is increased, and the unit’s size decreases the number of visible mortar joints, thus giving
large walls a different visual scale. The PCI Standard
contains the most of popular thin brick face sizes.
Contact precaster to determine availability of desired
color or texture in the face sizes selected.

(c)

(d)

Edge Corner

Close tolerances also can be obtained by saw-cutting
each brick, but this increases costs substantially.
FBX brick may be split into soaps (half brick). Often only
one side of the brick can be used as the facing veneer.
The use of soaps will increase the thickness and weight
of the panel. Whole bricks are not recommended for use
in precast concrete because of the difficulty in adequately
filling the mortar joints and the potential for freezingand-thawing spalling.
Figures 3.5.51 through 3.5.60 illustrate various projects
with applications of brick-faced precast concrete panels.

Some bricks (TBS or FBS, for example) are too dimensionally inaccurate for applications with precast
concrete panels. These bricks typically have high absorption rates that cause greater chances of efflorescing and freezing-and-thawing spalling. They conform
to an ASTM specification suitable for site laid-up applications, but they are not manufactured accurately
enough to permit their use in a preformed grid that
positions bricks for a precast concrete panel. Tolerances
in an individual TBX or FBX brick of ± 5/32 in. (±4 mm) or
more cause problems for the precast concrete producer. Brick (TBX and FBX) are available from some suppliers to the close tolerances necessary for precasting.
FBS and FBX are designations for facing brick types
that control tolerance, chippage, and distortion. Type
FBS is brick for general use in masonry while Type FBX is
brick for general use in masonry where a higher degree
of precision and lower permissible variation in size than
permitted for Type FBS is required (see ASTM C 216).
For thin-veneer brick units, Type TBS (Standard) is thinveneer brick for general use in masonry while Type TBX
(Select) is thin-veneer brick for general use in masonry
where a higher degree of precision and lower permissible variation in size than permitted for Type TBS is required (see ASTM C 1088).

End Edge Corner

(b)

(c)

(a)
Fig. 3.5.51(a), (b) & (c)
San Francisco Museum of Modern Art, San Francisco, California;
Architect: Mario Botta, Design Architect; Hellmuth, Obata &
Kassabaum, P.C. (HOK) Architect of Record; Photos: Perretti &
Park Pictures.

ARCHITECTURAL PRECAST CONCRETE

|

191

3

SURFACE AESTHETICS
3.5.10.4 Clay product selection

The patterned façade on the museum
(Fig. 3.5.51) is composed of bands of rusticated red brick accented by flamed white
and black granite on the upper level. The
1-in.-thick (25 mm) bricks are cast in 9-in.thick (225 mm) precast concrete panels.
The bricks, some 600,000 in all, were rolled
in sand before baking to give them a grainy
finish. Most panels measure 10 × 281/2 ft (3
× 8.7 m) and contain 1500 to 2300 bricks
per panel. The museum’s horizontality was
emphasized by raking the mortar joints between brick courses (Fig. 3.5.51[b] and [c]).
These figures also show the flat and corner panels with corner brick, as well as the
close-ups of the façade patterns.

(b)

(c)

Fig. 3.5.52(a), (b) & (c)
Merrill Lynch Hopewell Campus
Pennington, New Jersey;
Architect: Thompson, Venulett,
Stainback & Associates (TVS);
Photos: Brian Gassel/TVS.

192

|

ARCHITECTURAL PRECAST CONCRETE

(a)

SURFACE AESTHETICS

3.5.10.4 Clay product selection

The eight office and four assembly buildings in the
1.5 million ft2 (139,300 m2) campus shown in Fig.
5.3.52(a) are clad with 6882 clay product–faced architectural precast concrete panels totaling 548,623 ft2
(50,967 m2). The precast concrete panels embedded
with thin brick provided a basic kit of parts. The kit of
parts did not just rely on basic columns and spandrels
common in office buildings but instead was based on
typical modules: a typical 30 ft (9 m) bay with a fourth
story variation, an entry bay, a gable end with curtain
wall, a gable end with precast concrete cladding, stair
towers, and arcades. The elements are combined in
different floor plan configurations and building heights
to achieve the diversity of sizes and shapes of the required facilities. The ability to combine the brick and
the trim in a single panel using thin bricks and buffcolored, sandblasted precast concrete bands was a key
element in the building design. The typical bay was designed to have horizontal joints visible only where brick
relief joints would normally be located and the vertical spandrel joints are located behind recessed downspouts (Fig. 3.5.52[b]). The panels also clad four parking structures and the thin-brick panels were brought
into the dramatic main dining room in the assembly
buildings (Fig. 3.5.52[c]). With multiple buildings under
construction simultaneously, having brick-clad precast
concrete panels produced off-site helped to reduce the
on-site work required. That helped reduce the amount
of people, equipment, and materials on the job and
ultimately created a more manageable, cleaner, and
safer worksite.
Fig. 3.5.53
Nordstrom Palm Beach Gardens; Palm Beach Gardens, Florida;
Architect: Callison Architecture Inc.; Photo: Vern Smith.

3

Fig. 3.5.54
Centergy at Technology Square
Atlanta, Georgia;
Architect: Smallwood, Reynolds,
Stewart, Stewart & Associates Inc.;
Photo: Gabriel Benzur.

The store in Fig. 3.5.53 features insulated brick-faced
precast concrete panels highlighted with bands of 4
× 12 in. (100 × 300 mm) utility brick at the entry that
alternate with the precast concrete tones. The panels
were designed as shearwalls and have a light sandblast
finish on the accent stripes. Brick-faced precast concrete panels were selected because the job schedule
was able to be reduced by four months versus conventional masonry.
Brick-faced precast concrete panels were specified for the office complex in Fig. 5.3.54 over traditional brick construction for its cost efficiencies,
speed of construction, and simplified logistics. The
complex contains three structures: a 6-story office
building, a 14-story office tower, and an 8-story
parking structure that sits behind the two office
buildings. The bricks are 5/8 in. (16 mm) thick and
are the skin of 6-in.-thick (150 mm) concrete panels. Approximately 300 brick-faced precast concrete panels, including some as long as 40 ft (12.2
m) and weighing 15 ton (13.6 t) form the shell of
the complex. Designed as a nexus for a thriving
high-tech corridor, the project connects Georgia
Tech University with a burgeoning business and
residential community. Architectural precast concrete panels helped mix town and gown in a style
that fit both neighborhoods.

ARCHITECTURAL PRECAST CONCRETE

|

193

3

SURFACE AESTHETICS
3.5.10.4 Clay product selection

Fig. 3.5.55
Hull Street Parking Deck
Athens, Georgia;
Architect: Smallwood, Reynolds, Stewart, Stewart & Associates;
Photo: Jim Roof.

The 950-vehicle parking structure (Fig. 3.5.55) was
designed to address a university town’s acute parking
shortage while blending with the classical architecture of the campus buildings. An all-precast concrete
structure was selected due to aesthetics, economy, and
speed of construction. The fast-track schedule took

advantage of the ability to cast components, which included both structural and exterior façade components,
before the completed design package was issued. Inset
thin brick was used on upper-level panels, with the
panels cast with the brick in place in the molds, creating a one-step operation. Lower floors feature panels

(b)
(a)

Fig. 3.5.56
Woodmont High School, Piedmont, South Carolina; Architect: Perkins & Will, Design Architect; and Craig Gaulden Davis,
Architect of Record; Photos: Craig Gaulden Davis Architects/Photographer – Working Pictures.

194

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.10.4 Clay product selection

3

with a limestone-like appearance that was achieved
with a buff-colored finish, light sandblast texture, and
detailed reveals. This combination was accented by
tall, classic columns and arched windows at the stair
tower, which draw attention and create a dramatic
appearance. Only a few different sizes and shapes of
precast concrete panels were required, speeding production and reducing costs by minimizing the number
of molds. Material needs also were reduced by using
the exterior precast concrete as both the façade and as
loadbearing panels for the interior double tees.
The high school in Fig. 3.5.56(a) and (b) is a new
250,000 ft2 (23,000 m2) school with a core for 2000
students. The thin-set red brick on the exterior was
used to give the precast concrete a traditional feel that
appealed to the community. Sandblasted reveals between the brick panels organize the different levels of
the building and reduce the scale of the precast concrete panels. A stack bond masonry pattern was chosen
to minimize the visibility of joints between the panels.
Creating a segmented curve along the exterior wall of
the media center gave the effect of the school opening
its arms to the public and calling attention to the main
entrance (Fig. 3.5.56[a]). The insulated sandwich panels
of the media center allowed the interior finish to be
exposed painted precast concrete (Fig. 3.5.56[b]).
Design features of the seven-story medical office
building in Fig. 3.5.57 include embedded, multi-colored brown brick laid in a series of patterns—stacked,
running, Flemish bond, and soldier coursing—created
to minimize the scale of the building. The brickwork is
highlighted by blocks of acid-etched precast concrete
accents and bands, as well as curved column covers.

Fig. 3.5.57
Sparrow Professional Building
Lansing, Michigan;
Architect: Albert Kahn Associates Inc.;
Photo: Glen Calvin Moon/Albert Kahn Associates Inc.

The three-story building in Fig. 3.5.58 is clad with thin
brick–faced precast concrete panels that incorporate
precast concrete sills, jambs, and headers, as well as
banding in monolithic units. Radial brick were cast into
radiused precast concrete panels to create the smooth
flow around corners.

Fig. 3.5.58
S. C. Johnson Worldwide Headquarters, Mt. Pleasant, Wisconsin;
Architects: Zimmerman Design Group; and Hellmuth, Obata & Kassabaum, P.C.; Photo: Edward Purcell.

ARCHITECTURAL PRECAST CONCRETE

|

195

3

SURFACE AESTHETICS
3.5.10.4 Clay product selection

(b)

Fig. 3.5.59(a) & (b)
Garfield Heights High School Academic and Health and Physical
Education Buildings, Garfield Heights, Ohio;
Architect: Arcadis FPS;
Photos: Arcadis FPS, Cleveland.

Fig. 3.5.60
Miller Park, Milwaukee, Wisconsin;
Architect: HKS Sport & Entertainment Group;
NBBJ Sports & Entertainment; and Eppstein
Uhen Associated Architects (joint venture);
Photo: Eric Oxendorf.
(a)

The school project in Fig. 3.5.59(a) and (b) consists of
two buildings: a three-story academic building and a
one-story health and physical education building. Both
buildings are clad with 10-in.-thick (250 mm), thin
brick–faced, insulated precast concrete panels. The
panels consist of 5-in.-thick (125 mm) concrete back
wythe, 2 in. (50 mm) of insulation, and a 3 in. (75
mm) concrete face wythe embedded with thin brick.
The 5 in. (125 mm) back wythe was strong enough to
handle all precast concrete connections and lifting inserts without penetrating the insulation, ensuring the
maximum insulating value.
The 42,500-seat home of the Milwaukee Brewers is
a modern engineering marvel with traditional baseball flavor. Its unique, fan-shaped structure features
the first retractable roof of its kind in the world. The
building’s façade features radiused, brick-faced architectural precast concrete panels to convey the classic
look of past ballparks combined with wide, arched
windows reminiscent of historic European train stations (Fig. 3.5.60).

196

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.10.4 Clay product selection

(a)

3

Fig. 3.5.61(a) & (b)
The Paul Cejas School of Architecture,
Florida International University, Miami, Florida;
Architect: Bernard Tschumi Architects, and BEA
International (joint venture);
Photos: Thomas Delbeck.

(b)

Glazed and unglazed ceramic tile units should conform to American National Standards Institute (ANSI)
A 137.1, which includes American Society for Testing
and Materials (ASTM) test procedures and provides a
standardized system to describe the commonly available sizes and shapes, physical properties, basis for
acceptance, and methods of testing. Ceramic tiles are
typically 3/8 to 1/2 in. (10 to 13 mm) thick, with a 11/2%
tolerance on the length and width measurements.
When several sizes or sources of tile are used to produce a pattern on a panel, the tiles must be manufactured on a modular sizing system in order to have
joints of the same width.
Glazed units may craze from freezing and thawing
cycles or the bond of the glaze may fail due to exposure to extreme environmental conditions. The body
of a tile (not the glazed coating) must have a water
absorption of less than 3% (measured using ASTM C
373) to be suitable for exterior applications. However,
low water absorption alone is not sufficient to ensure
proper selection of exterior ceramic tiles. As a result,
when ceramic tile is required for exterior use, the manufacturer should be consulted for frost-resistant materials for exterior exposure. Glazes are covered by ASTM
C 126 and tested in accordance with ASTM C 67.

The gallery and lecture halls of the School of
Architecture in Fig. 3.5.61(a) and (b) are not traditional
precast concrete buildings, structurally or architecturally, although the walls are loadbearing and support
floor and roof double tees. The architectural expression
is colorful ceramic tile and a variety of outdoor spaces.
The architect sculpted a pair of engaging forms, then
wrapped them in red, orange, and yellow ceramic tile
that gives the ensemble a hot, Latin flair. The vivid yellow and red structures are clad with 8 × 8 in. (200 ×
200 mm) ceramic tiles with brilliant color variations.
Tiles were recessed into the precast concrete, which
produced a tightly sealed flush edge joint at the lightly
sandblasted panel borders.

ARCHITECTURAL PRECAST CONCRETE

|

197

3

SURFACE AESTHETICS
3.5.10.4 Clay product selection

(a)

(a)

Fig. 3.5.62(a) & (b)
Prospect Heights Care Center
Hackensack, New Jersey;
Architect: Herbert Beckhard Frank Richlan & Associates;
Photos: Norman McGrath Photograph.

Fig. 3.5.63
Saks Parking Garage, Kansas City, Missouri; Architect: Gastinger
Walker Harden Architects; Photo: Mike Sinclair.

198

|

ARCHITECTURAL PRECAST CONCRETE

The façades of the building in Fig. 3.5.62(a) and (b)
are sheathed in precast concrete from the ground up.
A number of panels are gull wing–shaped with wings
containing windows angling outward at 45° on each
end. These panels are 24 ft (7.5 m) long and 7 ft 3 in.
(7.8 m) high. The panels above ground level have 8 ×
8 in. (200 × 200 mm) brick-colored tile inserts, adding
a degree of contrast with the concrete while blending
harmoniously with the predominately brick neighboring buildings. The use of clay tiles inset within the precast concrete panels provides a greater variety of color
and texture than standard precast concrete panels. The
clay tiles feature keybacks around which the concrete
set, assuring permanent adherence. The end panels
were formed with concrete returns to avoid miters or
revealing actual panel thickness.
The mixed-use, five-level parking structure in Fig.
3.5.63 was designed to blend with the surrounding

SURFACE AESTHETICS

3.5.10.4 Clay product selection

3

Fig. 3.5.64(a) & (b)
88 Kearney Street
San Francisco, California; Architect: Skidmore, Owings and Merrill;
Photos: Skidmore, Owings and Merrill San Francisco.

(a)

structures with their highly detailed and ornamented
Spanish architecture. Areas of smooth texture on the
architectural precast concrete panels provided a base
for tile patterns to enliven the façade and create an
overall structure that serves to further enrich the area.
No ASTM standards exist for terra cotta, but units
should meet the minimum requirements published by
the Architectural Terra Cotta Institute. Architectural
terra cotta is a custom-made product and, within certain limitations, is produced in sizes for specific jobs.
Two thicknesses of terra cotta are usually manufactured: 1¼-in.-thick (32 mm) and 21/4 -in.-thick (56 mm)
units. Sizes range from 20 to 30 in. (500 to 760 mm)
for 11/4 in. units to 32 × 48 in. (810 × 1220 mm) for
21/4 in. units. Other sizes used are 4 or 6 ft × 2 ft (1.2
or 1.8 m × 0.6 m). Tolerances on length and width are
a maximum of ± 1/16 in. (+1.6 mm) with a warpage
tolerance on the exposed face (variation from a plane
surface) of not more than 0.005 in. (0.12 mm) per 1 in.
(25 mm) of length. The use of terra cotta–faced precast
concrete panels for restoration and new construction is
illustrated in Fig. 3.5.64, 3.5.65, and 3.5.66.

(b)

Built in 1906, the six-story building in Fig. 3.5.64(a)
and (b) is considered one of San Francisco’s architectural landmarks. For that reason, it was decided the
building’s terra cotta façade would be preserved on an
otherwise all-new structure of slightly taller height. The
terra cotta was taken off the building, piece by piece
and identified for subsequent reassembly on new precast concrete panels. Stainless steel wires were looped
through the back ribs of the terra cotta pieces and projected into the backup concrete to anchor the pieces
to the concrete.

ARCHITECTURAL PRECAST CONCRETE

|

199

3

SURFACE AESTHETICS

3.5.10.4 Clay product selection / 3.5.10.5 Design considerations

(a)

Fig. 3.5.65(a), (b) & (c)
Sacramento County Systems and Data Processing
Sacramento, California; Architect: HDR Architecture Inc. formerly
Ehrlich-Rominger; Photos: HDR Architecture Inc.

(b)

(c)

Precast concrete panels with 1-in.-thick (25 mm)
brick on 5-in.-thick (125 mm) concrete panels along
with glazed terra cotta on the spandrels and mullions
clad the nine-story building in Fig. 3.5.65(a). Panels of
light and deep sandblast finishes tied both systems together. See Fig. 3.5.65(b) and (c) for a close-up of the
terra cotta units.
For the sake of the traditional look of the historic
Michigan Avenue streetwall’s appearance, terra cotta–
faced precast concrete was used for the 260,000 ft2
(24,200 m2 ) retail/cinema building (Fig. 3.5.66[a]), encompassing an entire block. The terra cotta pieces are
a variety of shapes and sizes, with some flat, fluted, or
round (Fig. 3.5.66[b]). The backs of the extruded pieces were flat and holes were drilled in the terra cotta for
insertion of stainless steel pins. The terra cotta units
were placed in a mold and 10 in. (250 mm) of concrete
was then cast to create a panelized system.
Variations in brick or tile color will occur within and
between lots. The clay product supplier must preblend
any color variations and provide units that fall within
the color range specified and approved by the architect
for the project. Defects such as chips, spalls, face score

200

|

ARCHITECTURAL PRECAST CONCRETE

lines, and cracks are common with brick, and the defective units should be culled from the bulk of acceptable units by the clay product supplier according to the
architect’s requirements and in accordance with applicable ASTM specifications. Should minor damage occur to the clay product face during shipping, handling,
or erection, field remedial work can be accomplished,
including replacement of individual clay products. Units
may be chipped out and new units installed using an
epoxy, dry-set, or latex portland cement mortar.

3.5.10.5 Design considerations
The clay product surfaces are important in order to
bond to the backup concrete. Textures that offer a
good bonding surface include a:
• Scored finish, in which the surface is grooved
(ribbed) or dovetailed (keybacked) as it comes from
the die
• Combed finish, in which the surface is altered by
parallel scratches
• Roughened finish, which is produced by wire cutting or wire brushing to remove the smooth surface
or die skin from the extrusion process.
• A brick wire cut (through extruded holes in whole
bricks) to provide two (half) soaps.

SURFACE AESTHETICS

3.5.10.5 Design considerations

With thin- and half-brick units, no metal ties or weeps
are required to attach them to the concrete because
adequate bond is achieved. In general, clay products
that are cast integrally with the concrete have bond
strengths exceeding that obtained when laying units
in the conventional manner (clay product to mortar). In
pullout tests, the brick fails or shears before it pulls out
of the concrete. It is necessary, however, to be careful
to not entrap air or excess water-caused voids. These
voids could reduce the area of contact between the
units and the concrete, thereby reducing bond.

quarry tile and frost-resistant glazed wall tiles generally
do not need to be wetted. Terra cotta units should be
soaked in water for at least one hour prior to placement to reduce suction and they should be damp at
the time of concrete placement.

The bond between the clay product facing and the
concrete depends on the absorption of the clay product and the concrete’s water-cement ratio. Either low
or high absorption will result in a poor bond. Half
bricks with a water absorption of 6 to 9% obtained by
five-hour boiling provide good bonding potential. Thin
bricks should have a water absorption less than 6%
per the PCI Standard.

When removed from the kiln after firing, clay bricks
will begin to permanently increase in size as a result of
absorption of atmospheric moisture. The expansion of
the clay products can be absorbed by four simultaneously occurring negative dimensional changes of the
clay product and concrete:

Half bricks with an initial rate of absorption (suction)
of less than 30 g /30 in.2 per min (30 g /194 cm2 per
min), when tested in accordance with ASTM C 67, are
not required to be wetted. However, brick with high
suction or with an initial rate of absorption in excess of
30 g /30 in.2 per min should be wetted prior to placement of the concrete. This will reduce the amount of
mixture water absorbed and improve bond. Unglazed

2. Elastic deformation of the concrete under stress

3

Because of the differences in material properties between the facing and concrete, clay product–faced
concrete panels may be more susceptible to bowing
than homogeneous concrete units. However, panel
manufacturers have developed design and production
procedures to minimize bowing.

1. Drying shrinkage of the concrete

3. Creep of the concrete under stress
4. Elastic deformation of the clay product under
stress
In general, strains imposed slowly and evenly will not
cause problems. During the first six months to a year
after panel production, (Fig. 3.5.67), tile expansion is
(a)

(b)

Fig. 3.5.66(a) & (b)
600 North Michigan Avenue
Chicago, Illinois;
Architect: Beyer Blinder Belle, Design
Architect; and Shaw and Associates,
Architect of Record.

ARCHITECTURAL PRECAST CONCRETE

|

201

3

SURFACE AESTHETICS
3.5.10.5 Design considerations

1/4”

Concrete

1 1/8”

Perimeter
Reveal

Brick

Concrete
Coved
Mortar
Joints
Brick
Concrete Panel Section
Mortar Joint and Reveal Details

Fig. 3.5.67 Relative temperature and moisture movements of concrete, brick, tile
and mortar.

Fig. 3.5.69 Mortar joint and reveal details.

Fig. 3.5.68 Corner details.

3 5/8”
Concrete
Panel

Caulking color
to match mortar
joints
Concrete
Panel

7 5/8”
3 5/8”

Caulking
color to
match brick

|

3/8”

24”
Brick

Corner brick
7 5/8” x 3 5/8” x 2 1/4”

4”
4”

3 5/8” 3/8”

Brick

(a) Indented Corner Detail

202

7 5/8”

ARCHITECTURAL PRECAST CONCRETE

(b) Wrapped Corner Detail

(c) Deep Return

SURFACE AESTHETICS

3.5.10.5 Design considerations / 3.5.10.6 Production and construction considerations

small and the rate of strain application is slow, but concrete shrinkage is nearly complete. The concrete creeps
under the load to relieve the tensile stress generated
in the tile when the concrete shrinks because the tiles
are relatively rigid (elastic modulus of tile/elastic modulus of concrete). After this time period, the tile has
many years to accommodate the additional moisture
expansion.
Three types of corner details may be used: (1) indented (Fig. 3.5.68[a]); (2) wrapped (Fig. 3.5.68[b]); or (3)
deep return (Fig. 3.5.68[c]). Brick mortar joints should
be concave (cove). At reveals and at the top and bottom of inset areas, the concrete should cover the edges
of the brick units (Fig. 3.5.69). The designer also needs
to pay special attention to where the joints between
concrete panels are located, which is a departure from
the use of traditional brick masonry.

3

In the development of the technical documents, using
thin brick precast concrete panels provides a simplification of detailing over hand set masonry. The system
avoids intricate flashing, masonry support, and masonry
anchoring requirements that would be necessary with
conventional construction to achieve layering and relief
features.

3.5.10.6 Production and construction
considerations
Clay product–faced units have joint widths controlled
by locating the units in a suitable template or grid system set out accurately on the mold face (Fig. 3.5.70).
Common grid systems generally consist of an elastomeric (or rubber) form liner or a plastic form liner. Liner
ridges are typically shaped so that joints between units
simulate concave-tooled joints.
Tolerances for brick-faced precast concrete panels are
shown in Fig 3.5.71. The number of bricks that could
exhibit any misalignments should be limited to 2% of
the bricks on the panel.
Tiles, measuring 2 × 2 in. (50 × 50 mm) or 4 × 2 in.
(100 × 50 mm), may be supplied face-mounted on
polyethylene or paper sheets and secured to the mold
by means of double-faced tape or a special adhesive.

Fig. 3.5.70

Fig. 3.5.71 Tolerances for brick-faced architectural elements.

A

d

e
b
a

b
e

A-A

c

a

A
d

a = Alignment of mortar joints:
Jog in alignment………………………………………………………………………………… 1/8 in. (3 mm)
Alignment with panel centerline… ……………………………………………………………… ±1/8 in. (±3 mm)
b = Variation in width of exposed mortar joints……………………………………………………… ±1/8 in. (±3 mm)
c = Tipping of individual bricks from the panel plane of exposed brick surface… …………………… -1/4 in. (-6 mm), ≤ depth of form liner joint
d = Exposed brick surface parallel to primary control surface of panel… …………………………… +1/4 in., -1/8 in. (+6 mm, -3 mm)
e = Individual brick step in face from panel plane of exposed brick surface… ……………………… -1/4 in. (-6 mm), ≤ depth of form liner joint

ARCHITECTURAL PRECAST CONCRETE

|

203

3

SURFACE AESTHETICS

3.5.10.6 Production and construction considerations / 3.3.10.7 Application of clay products after casting of panel

Fig. 3.5.72
The Nikko Hotel
San Francisco, California;
Architect: Patri-Merker
Architects formerly
Whisler-Patri;
Photo: Patri-Merker
Architects.

The space between the tiles is filled with a thin grout
and then the backup concrete is placed prior to initial
set of the grout. Figure 3.5.72 shows a project that uses
2 × 2 in. (50 × 50 mm) tiles that have been placed with
the method described. For the best appearance, narrow
tile joints should be filled from the front, particularly if
cushion-edged tiles are used.
Fig. 3.5.73 The anchorage of brick to precast concrete using
dovetail anchor and slot.
4" brick wall

Dovetail
slot
Flexible
dovetail
anchor
PLAN
3/16" Dia.
(W2.8)
3" Min.

|

Face of concrete

7/8"
No. 12 Ga.

204

Brick face
of brick
veneer

SECTION

1" Max. air space

ARCHITECTURAL PRECAST CONCRETE

No. 22 Ga. steel
dovetail slot

3.5.10.7 Application of clay products
after casting of panel
Full brick may be installed on an already-cast precast
concrete panel at the plant or at the jobsite. Bricks
generally bear on a ledge created by a recess on the
precast concrete panel surface or on a shelf angle.
Full brick, supported on concrete ledge or steel shelf
angle, requires anchors for lateral support. The anchors
should be flexible and capable of resisting tension and
compression forces perpendicular to plane of the wall,
but permitting slight vertical and horizontal movement parallel to the plane of the wall. This flexibility
allows differential movements between the precast
concrete and the clay product veneer without cracking
or distress.
Galvanized or stainless steel wire anchors (ASTM A 82
or B 227, Grade 30HS) should be at least 3/16 in. (W 2.8)
in diameter and hooked on one end and looped through
a 7/8-in.-wide (22 mm), 12-gage steel sheet bent over
the wire (Fig. 3.5.73). The steel sheet is dovetailed on
the other end to fit into minimum 22-gage dovetail slot
in the concrete panel. The dovetail adjusts vertically so
the wire anchor can be placed in the bed joint of the
brick.
It has been found that a 16-gage dovetail anchor slot
fails at approximately the same load as a 26-gage slot
embedded in concrete, so there is not much advantage
to using heavier anchor slots to achieve increased load
capacity. Instead, more anchors should be used to obtain the required load capacity.
The minimum 3-in.-wide (75 mm) wire anchors
should be embedded at least 11/2 in. (38 mm), preferably 2 in. (50 mm), into the bed joint of the brick, with a
minimum 5/8 in. (16 mm) cover of mortar between the
anchor and the exterior wall face. The size and spacing
of anchors are based on tensile and compressive loads
induced by wind suction and pressure on the walls.
Most designers use the simple force multiplied by the
contributory area to determine anchor loads. When
this technique is used, additional anchors should be
provided at all openings and discontinuities, such as
windows, shelf angles, and concrete ledges, where
stresses are known to be higher. They should be
spaced not more than 3 ft (0.9 m) apart around the
perimeter of an opening and within 12 in. (300 mm)
of the opening.

SURFACE AESTHETICS

3.3.10.7 Application of clay products after casting of panel

Fig. 3.5.74
Oriole Park at Camden Yards
Baltimore, Maryland
Architect: Hellmuth, Obata and Kassabaum
Sports Facilities Group; Photo: HOK Sport.

There should be one anchor per 41/2 ft2 (0.42 m2) of
wall area. The maximum spacing between anchors
should not exceed 24 in. (600 mm) vertically and 36
in. (910 mm) horizontally. Anchors in alternate courses
should be staggered. Applicable building codes should
be consulted for additional reinforcement requirements, such as those for resistance to seismic forces
acting parallel to panels and for stack bond (which is
weaker than running bond).
The published tests on dovetail anchor slots and
dovetail anchors indicate an ultimate tension range of
713 to 965 lb (323 to 438 kg) and ultimate compression with a 1 in. (25 mm) cavity of 560 lb (254 kg) for a
12-gage dovetail anchor in a 22-gage dovetail slot (Fig.
3.5.73). A safety factor of 3, based on failure mode,
should be applied to arrive at design values.
To avoid anchor buckling, the distance between the
inside brick face and the concrete panel should not be
greater than 1 in. (25 mm) or less than 1/2 in. (13 mm).
This space should be kept free of mortar or other rigid
material to permit the differential movement between
the concrete panel and brick.
The low-scaled, arched façade in Fig. 3.5.74 presents
itself in the form of elaborate cornices and rustication
joints. Brick anchored to dovetail slots in the precast
concrete was field-laid on concrete ledges via scaffolding following precast concrete erection.

3

veneer at each floor, or at least every other floor, in
place of a concrete ledge (Fig. 3.5.75). The shelf should
be made of structural steel conforming to ASTM A 36
and properly sized and anchored to carry the imposed
loads. Anchor bolt holes should be horizontally slotted
to allow for ease of construction and horizontal movement. For shelf angles supporting unreinforced masonry, deflection should be limited to L/600, but should
not exceed 0.3 in. (7.5 mm). A small space should be
left between the lengths of angles to allow for horizontal thermal movements.
For severe climates and exposures, consideration
should be given to the use of galvanized or stainless
steel shelf angles. When using shelf angles, continuous
flashing should be installed over the angle. To ensure
adequate resistance to corrosion, coatings or materials
should conform to ASTM A 123 or ASTM A 167. The
suggested minimum level or corrosion protection for
coatings of anchor material is either ASTM A 153 Class
B-2 or ASTM A 167 Type 304.
Horizontal pressure-relieving joints should be placed
immediately beneath each shelf angle. Pressure-relieving joints may be constructed by either leaving an air
space or placing a highly compressible material under
the shelf angle and sealing the joint with an elastic
sealant and backer rod (Fig. 3.5.75).
Flashing in a masonry wall supported on shelf angles
is important for the proper drainage of water that may
penetrate the masonry. Flashing is not required if the
masonry is supported on a concrete ledge that has a
slope of 1/8 in. (3 mm) in 5 in. (125 mm), although

Fig. 3.5.75 Shelf angle with flashing and weep holes.

Flashing installed in
continuous reglet

Weep holes
Flashing extension
Elastic Sealant and backer rod
Compressible joint material
1/8 in. (3 mm) minimum
space below shelf angle

Shelf angles may be used to support the full-brick

ARCHITECTURAL PRECAST CONCRETE

|

205

3

SURFACE AESTHETICS

3.3.10.7 Application of clay products after casting of panel / 3.5.11 Honed or Polished

weep holes are necessary. Flashing materials are generally formed from sheet metals, bituminous-coated
membranes, plastics, vinyl, or combinations thereof,
the selection being largely determined by cost and suitability. The cost of flashing materials varies widely. Only
superior-quality materials should be selected, however,
since replacement in the event of failure would be exceedingly expensive.
Flashing may be installed in a continuous reglet or
recess in the concrete. To be most effective, the flashing should extend 1/2 in. (13 mm) beyond the wall surface and be turned down at a 45° angle to form a
drip. Weep holes, at least 1/4 in. (6 mm) in diameter,
should be provided in head joints immediately above
the flashing or concrete ledge at intervals of 16 to 24
in. (400 to 600 mm) maximum to permit drainage of
accumulated water.
If, for aesthetic reasons, it is necessary to conceal
the flashing, the number and spacing of weep holes
are even more important. In these cases, the spac-

ing should not exceed 16 in. (400 mm) on center.
Concealed flashing with tooled mortar joints can retain
water in the wall for long periods of time, thus concentrating the moisture at one spot.

3.5.11 Honed or Polished
Grinding of concrete removes the thin layer of cement
paste and cuts the aggregates to a uniformly smooth
surface. The grinding is called honing or polishing, depending on the degree of smoothness of the finish. The
surface can be a dull matte finish (honed) or, with the
use of increasingly finer grinding pads, can reach a high
luster (polished), depending also on the type of aggregates. Polished, exposed-aggregate concrete finishes
compare favorably with polished natural stone façades,
yet offer the architect total freedom of design while using the full structural capability of concrete. Honed and
polished finishes have gained acceptance because of
their appearance and excellent weathering characteristics, making them ideal for high traffic areas and polluted environments. Polished surfaces will also reflect more
heat than other finishes. Because of a corrosive and dirtladen atmosphere, a dense, polished surface texture
was used on the building in Fig.3.5.76. Maintenance
since 1973 has proven to be minimal for the architectural precast concrete cladding. The panels were made
with an exposed-quartz aggregate, silica sand, and
buff-tinted white cement. After fabrication and removal
from the mold, the panels were ground and polished
smooth on all flat surfaces. Figure 3.5.76 shows the
excellent details. While these are among the most expensive precast concrete finishes, they usually cost far
less than dimension stone.
There are only a few North
American producers with
equipment to produce the
polished finish.
In order to produce a
good ground or polished
finish it is first necessary
to produce a good plain
finish. The compressive
strength of the concrete
Fig. 3.5.76 Blue Cross/Blue
Shield Service, Center; Detroit,
Michigan; Architect: Giffels, Inc.
formerly Giffels & Rossetti, Inc.;
Photos: Daniel Bartushs, Giffels
Inc. – Southfield Office.

206

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.11 Honed or Polished

should be 5000 psi (34.5 MPa) before the start of
any honing or polishing operations. All patches and
the fill material in any bug/blow holes or other surface blemishes must also be allowed to reach approximately 5000 psi (34.5 MPa). It is preferable that the
matrix strength of the concrete mixture approach the
compressive strength of the aggregates or the surface
may not grind evenly or polish smoothly, producing
dull patches and possible dislodgement of aggregate
particles.
Because aggregates will polish better than the matrix,
it is essential to have a minimum matrix area. Either a
continuous or gap-graded concrete mixture carefully
designed to provide maximum aggregate density on
the surface to be polished is acceptable. In choosing
aggregates, special attention should be given to maximum size and hardness. Not all aggregates will accept
polish. Softer aggregates such as marble or onyx are
much easier to polish than either granite or quartz,
although the latter are preferred for their potentially
high polish. Limestone and quartzites will not accept a
polish while basalt can be honed to produce a smooth
but matte finish, but cannot be highly polished. Marble
aggregates may not retain their polish over time due to
chemical reaction with atmospheric pollutants, such as
acid rain. The hardness of the aggregates will affect
the rate of wear and tear on the grinding pads, which
in turn will affect the cost.
Of major concern to the economic viability of a polished project is the shape and geometry of the units
to be polished, the extent of polished finished face,
and the degree of hand polishing to edges, arrises, and
surface areas inaccessible to automatic machine heads.
Careful detailing to maximize use of automatic polishing equipment and minimize hand polishing will ensure
minimum cost. For reasons of economy, only surfaces
that can be ground completely by machines passing
over flat areas on returns should be honed or polished.
To a large degree, the amount of hand polishing is
governed by the shape of the panel. Hand polishing
of arrises, returns and other architectural features not
accessible with automatic equipment is slow and costly
and should not be designed, where possible.
Special detailing considerations that need to be noted
(Fig. 3.5.77) include:
• Flat surfaces are most easily polished; projections
should be avoided as they necessitate costly hand
work. Computerized polishing machines will polish

3

across grooves or recesses without difficulty, leaving
the surfaces within the grooves unpolished to provide contrast in color and texture. Alternatively, to
reveal the color of the aggregates, these grooves can
be sandblasted or acid-etched prior to polishing.
• Re-entrant angles should be avoided as they involved
expensive hand polishing. When a 90° return of a
panel is honed or polished, it may prove beneficial
to sequentially cast the return in a horizontal position. This will help create a more dense, uniform
surface. However, in the case of an element that is
L-shaped, all surfaces external to the re-entrant can
be polished.
• Convex surfaces with a radius of 10 ft (3 m) or more
can be polished by computerized polishing machinery but concave surfaces are unsuitable. Also, vertical and oblique planes ±45° to horizontal are capable of being polished without adjustment to the
element being polished. Columns of circular crosssection ranging between 8 to 50 in. (200 to 1200
mm) in diameter can be polished.
• Square, sharp edges should be avoided as they are
prone to chipping during polishing and damage
during handling. Miter joints should have a quirk
on the exterior corner.
• To reduce the risk of damage, both while polishing
and handling, rather than create a right angle at
the corner of a panel, an edge can be beveled to
produce a slightly obtuse angle.
• Alternately, the edge can be chamfered either by 3/8
to 1/2 in. (10 to 13 mm). Chamfers may be expensive to polish and should be left untreated if they
cannot be seen closely, or they may be sandblasted
prior to polishing.
• On panels incorporating more than one surface finish, the surface to be polished should be higher than
the other surfaces or separated by a wide groove.
The grinding process, which can be either wet or dry,
removes approximately 1/8 in. (3 mm) off the form face
of the precast concrete panels. Wet grinding is preferred, because the paste that is created aids in the
grinding. A high standard of craftsmanship is mandatory for this treatment, as the removal of the cement
skin emphasizes any defects in either formwork or compaction. It is very important to avoid any segregation in
the concrete. As with other finishes, final appearance
and uniformity will benefit if it is possible to match or
complement matrix color with aggregate color.

ARCHITECTURAL PRECAST CONCRETE

|

207

3

SURFACE AESTHETICS
3.5.11 Honed or Polished

Fig. 3.5.77 Detailing considerations.

Unsuitable for polishing

Suitable for polishing

Would require handwork
unpolished

FLAT SURFACES
• Requires hand work if returns are to be
polished.
Unsuitable for polishing
Unpolished

Taper:

Suitable for polishing

1
4
CURVED SURFACES

• Suitable – groove not polished. Groove may
be left as cast, acid etched, or sandblasted.

Unsuitable
Sharp edge

Suitable alternatives
Chamfer
Slight bevel

• Note need for taper or ‘draft’ to allow stripping from mold.
• Need for groove of adequate size, capable
of being ‘read’ at a distance.
Off-form recess

• Consider depth of groove in relation to
cover to reinforcement.

3/4” groove
Polished

Second
finish, eq.
sanblasted

To clear the
polishing head

Chamfer may also be
honed or
polished

CORNERS
Unsuitable details	Preferred details
Chamfer (or bevel) unpolished

• When more than one finish is required, the
surface to be polished should be ‘proud’ of
other surfaces.
Quirk
unpolished

JOINTS

Continued mechanical abrasion with progressively
finer grit, followed by filling of surface air voids and
rubbing, will produce a highly polished surface. The
depth of grinding determines the extent to which the
aggregate shows, but the color of the cement will be
important in any case. Such panels have an attractive
sheen that enhances many colors. Polished panels of

208

|

ARCHITECTURAL PRECAST CONCRETE

Chamfers
(or bevels)
unpolished

pastel colors tend to appear white when viewed from
a distance because of their high surface reflectance.
Therefore, this type of surface is recommended for
panels situated relatively close to the pedestrian traffic
flow or for those of medium or dark shades.
As is typical of a department store anchoring a subur-

SURFACE AESTHETICS

3.5.11 Honed or Polished

3

Fig. 3.5.78
Macy’s Department Store (Mall of America),Bloomington, Minnesota; Architect: Slomanson, Smith & Barresi Architects;
Photo: Slomanson, Smith & Barresi Architects.

(a)

ban shopping mall, the broad expanse of the exterior walls are windowless. The designer used surface
texture, the play of light and shade on receding wall
planes, strong entrance elements embracing the
compelling forms of the arch and the loggia in order to achieve a vibrant architectural composition
(Fig. 3.5.78). The 18-ft-high (5.5 m) panels from
the first to second levels combine two distinct color
mixtures and two surface finishes. The darker polished surfaces of quartzite and granite aggregates
were used to accent the plinth and the 4 ft (1.2
m) band courses, the entrances, central arches, and
loggias as a facing to frame the entrance porticos.
The lighter surfaces of quartzite aggregates was
given an acid-etch finish.
The panels on the embassy in Fig. 3.5.79(a) have
two different finishes using the same face mixture: a
very shiny, highly polished darker surface and a bushhammered, rougher, lighter surface. They form horizontal bands throughout the façades of the 13-story
Chancellery building and 9-story public building (Fig.
3.5.79 [b]). The darker polished surfaces simulate a
Mexican natural granite.
Fig. 3.5.79 (a) & (b)
French Embassy, Mexico, D.F., Mexico;
Architect: Bernard Kohn Associates; and
Eduardo Terrazas y Asociados (joint venture);
Photos: Bernard Kohn Associates.

(b)

ARCHITECTURAL PRECAST CONCRETE

|

209

3

SURFACE AESTHETICS
3.5.11 Honed or Polished

Above the base of the building in Fig. 3.5.80(a), the
granite used at the base was matched with a concrete
mixture with local, crushed-granite aggregate by honing the double window precast concrete panels. The
look of rusticated granite in a running bond pattern
was developed by 1 × 1 in. (25 × 25 mm) reveals that
were lightly sandblasted before the panel was honed
(Fig. 3.5.80[b]).
A combination of acid etching of the matrix area
with polishing of the aggregate produces a surface
characterized by flat, coarse aggregate that is slightly
proud of the underlying matrix. This surface is highly
resistant to weathering and is self-cleaning. A polished/sandblasted finish provides a contrast between
the polished aggregate and the sandblasted matrix of
the concrete.

tion is detailed to separate them—usually a V-groove.
The exact shape of this V-groove will require consultation with and trials by the precaster to ensure that the
angle within the V-groove is sufficiently flat to prevent
chipping the edge in the grinding process and causing
an unsightly line on one side of the groove.
The designer will also need to consider that the polishing process will create a step in the surface between
the polished and the acid-etched or sandblasted finish
of approximately 1/8 in. (3 mm), which can cause problems when aligning precast concrete units on site. This
can be overcome during the design process. Casting a
series of trial panels is strongly recommended to provide full knowledge of the combined effects.

(a)

The use of combination finishes requires the designer
to make an early decision to ensure that the overall
concept makes allowance for the change in color and
texture of the two finishes and that a suitable demarca(b)

Fig. 3.5.80(a) & (b)
BC Hydro Office Building
Vancouver, B.C., Canada;
Architect: Musson Cattell Mackey Partnership;
Photos: Simon Scott.

210

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.12 Stone Veneer-Faced Precast Concrete / 3.5.12.1 General considerations

3.5.12 S tone Veneer–Faced Precast
Concrete
Natural stone veneer–faced precast concrete has become widely used in building construction because of
its strength, durability, aesthetic effect, availability, and
inherent low-maintenance costs. The incorporation of
the stone veneer into the precast concrete panels provides an economically viable solution to cladding today’s structures.
Stone veneer–faced precast concrete panels offer
many benefits. These include:
1. Veneer stock can be used in thinner sections because anchoring points may be placed closer
together.
2. Multiplane units such as column covers, spandrels
with integral soffit and sill sections, deep reveal
window frames, inside and outside corners, projections and setbacks, and parapet sections are more
economically assembled as veneer units on precast
concrete panels. Often, it is desirable to use one of
the veneer materials in a traditional manner around
the lower portion of a building and extend a similar finish with veneered precast concrete panels up
the exterior walls.
3. Precast concrete systems permit faster enclosure,
allowing earlier work by other trades and subsequent earlier occupancy, because each of the larger panels incorporates a number of veneer pieces.
4. Veneered precast concrete panels can be used to
span column-to-column and floor-to-floor, thereby
reducing floor-edge loading and eliminating elaborate temporary scaffolding.

3.5.12.1 G
 eneral considerations
The purchaser of the stone should appoint a qualified
individual to be responsible for coordination, which
includes delivery and scheduling responsibility and ensuring color uniformity. Color control or blending for
uniformity should take place at the stone fabricator’s
plant, because ranges of color and shade, finishes,
and markings such as veining, seams, and inclusions
are easily seen during the finishing stages. Acceptable
stone color should be judged for an entire building elevation rather than as individual panels. The responsibility for stone coordination should be written into the
specifications so it can be priced. The owner, architect,
and/or stone purchaser should visit the stone fabri-

3

cator’s plant to view the stone veneer and establish
criteria and methods for color-range blending on the
project.
All testing to determine the physical properties of the
stone veneer-with the same thickness and finish that
will be used on the structure-should be conducted by
the owner prior to the award of the precast concrete
contract. This test data will determine the stone design
and anchor placement.
There is a need for close coordination between the
precast concrete manufacturer and stone veneer supplier. Shop drawing preparation and submissions may
vary from procedures established for non-veneered precast concrete panels. Checking and approval of these
details and shop drawings will be simplified and expedited if they can be combined and/or submitted simultaneously. Because of schedule issues, separate subcontracts and advance awards often occur in projects with
stone-veneered panels. While these procedures may affect normal submission routines, it is not intended that
responsibilities for accuracy should be transferred or reassigned. Typically, the precaster is responsible for precast concrete and stone layouts and details, while the
stone-veneer fabricator is responsible for stone shopfabrication drawings and drilling of anchor holes.
The production of stone veneer panels requires adequate lead time in order to avoid construction delays.
Therefore, it is important that approvals for shop drawings are obtained expeditiously. Furthermore, it is recommended that the designer allow the submission of
shop drawings in predetermined stages so production
can begin as soon as possible and ensure there is a
steady and timely flow of approved information to allow uninterrupted fabrication.
The precast concrete producer must provide the
stone quantity and sequence requirements to meet the
panel fabrication schedule. For reasons of production
efficiency, some concrete panels may be produced out
of sequence relative to the erection sequence. The precaster and stone fabricator should coordinate packaging requirements to minimize handling and breakage.
Extra stone (approximately 2 to 5%) should be supplied to the precaster to allow immediate replacement
of damaged stone pieces. The extra stone should be
the largest pieces to be used on the project. Deliveries
should be scheduled to accommodate actual panel
fabrication schedules.

ARCHITECTURAL PRECAST CONCRETE

|

211

3

SURFACE AESTHETICS
3.5.12.2 Stone properties

3.5.12.2 S tone properties
Stone is a product of geologic evolution and, therefore,
does not demonstrate the consistent behavior that may
apply to manufactured building materials, such as concrete. The strength of natural stone depends on several
factors: the size, rift, and cleavage of crystals; the degree of cohesion; the interlocking geometry of crystals;
the nature of natural cementing materials present; and
the type of crystal. The stone’s properties will also vary
with the locality from which it is quarried. Therefore, it
is important that current testing is performed on stone
quarried for each specific project.
Sedimentary and metamorphic rocks, such as limestone and marble, will exhibit different strengths when
measured parallel and perpendicular to their original bedding planes (anisotropic). Igneous rocks, such
as granite, may or may not exhibit relatively uniform
strength characteristics in their various planes (isotropic). In addition, the surface finish, freezing and thawing, and large temperature fluctuations will affect the
stone strength and in turn influence the anchorage
system required for the stone to the precast concrete.
Information on the durability of the specified stone
should be obtained through current testing in conjunction with observations of existing installations of that
particular stone. This information should include such
factors as tendency to warp, reaction to weathering
forces, resistance to chemical pollutants, resistance to
chemical reaction from adjacent materials, and reduction in strength from the effects of weathering or wetting and drying.

212

|

a measurable degree. Bushhammered and other similar
surface finishes also reduce the effective thickness and
strength. For 11/4 in. thick (3 cm) veneers, a reduction in
thickness of 1/8 in. (3 mm) reduces the theoretical bending strength by about 20% and increases the elastic
deflection under wind loads by about 37%.
Laboratory tests on 11/4 in. (3 cm) -thick specimens of
unaged, thermally finished granite revealed that the effects of the thermal finish reduced the bending strength
of the specimens by as much as 25 to 30%. The loss
of strength depends mainly on the physical properties
of the stone forming minerals, on the coherence of the
crystalline structure of the stone, and on the presence
of micro and macrofractures in the stone.
Thermal or flame finishing, and to a certain degree
bushhammering, of granite surfaces causes microfracturing, particularly of quartz and feldspars. These microcracks permit absorption of water to a depth of about 1/4
in. (6 mm) in the distressed surface region of the stone,
which can result in degradation by cyclic freezing and
thawing and a further reduction in bending strength.
Weathering affects different stones in different ways.
It can cause both a chemical decomposition and physical disintegration in some stones. The thinner the stone
is sliced, the more susceptible it may be to weathering.
Most natural stones lose strength as a result of aging
(thermal cycling, for example, heating to 170 °F [77
°C] and cooling to -10 °F [-23 °C], and wet/dry cycling).
The modulus of rupture of building stone can also be
affected by freezing and thawing of the stone.

Tests should be performed by the stone fabricator
to determine the physical properties of the stone being considered prior to awarding the precast concrete
contract. The testing should be done on stone with
the same finish and thickness that will be used on the
structure. Flexural tests (ASTM C 880) should be used
to evaluate the physical properties and obtain design
values. Absorption testing (ASTM C 97) helps evaluate freezing and thawing durability. These properties,
along with the performance of the anchors attaching
the stone veneer, should be used to ensure adequate
strength of the panel to resist loads during handling,
transportation, erection, and in-service conditions.

Flexural tests (ASTM C 880) should be conducted on
the selected stone, at the thickness and surface finish
to be used, in both the new condition and the condition after 100 cycles of laboratory-accelerated aging (weathering) tests to determine the reduction in
strength, if any. Suggested weathering test procedures
include cycling between 170 °F (77 °C) and -10 °F (-23
°C), while the face of the stone is submerged in a 4 pH
sulfurous acid solution that simulates chemical weathering. For warm climates, the test procedure can be
modified to cycle between 40 °F (5 °C) and 170 °F (77
°C). Also, in areas where the pH of rainfall is above 6,
the acid solution can be eliminated. Absorption testing
(ASTM C 97), as mentioned, helps evaluate the freezing and thawing durability of the stone.

The process used to obtain a thermal or flame finish on
granite veneers reduces the effective stone thickness by
about 1/8 in. (3 mm), as well as the physical strength to

Stones that have a satisfactory performance record in
thicknesses, sizes, and climates similar to those envisioned for a project may, at the option of the designer,

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.12.2 Stone properties / 3.5.12.3 Stone sizes

be exempt from the above testing requirements.
For most types of stone, temperature-induced movements are theoretically reversible. However, certain
stones, particularly marble, when subjected to a large
number of thermal cycles, develop an irreversible expansion in the material amounting to as much as 20%
of the total original thermal expansion. This residual
growth is caused by breaking of crystal bonds. Such
growth, if not considered in the stone size, may result
in curling or bowing of thin marble. For relatively thick
marble veneers, the expansion effects are restrained or
accommodated by the unaffected portion of the veneer. Tests should be performed to establish the minimum thickness required to obtain satisfactory serviceability. Stone can be exposed to differential accelerated
heating and cooling cycles and measured for deformation (bowing/hysteresis).
Table 3.5.2 Permeability of commercial building stones,
cu in./ft2/hr for ½ in. thickness.

Water Pressure, psi
Stone Type

1.2

50

100

Granite

0.06–0.08

0.11

0.28

Limestone

0.36–2.24

4.2–44.8

0.9–109

Marble

0.06–0.35

1.3–16.8 0.9–28.0

Sandstone
Slate

4.2–174.0

51.2

221

0.006–0.008 0.08–0.11

0.11

Note: 1 cu in./ ft2/hr/ 1/2 in. = 16.39 m3/hr/13 mm;
1 psi = 0.006895 MPa; 1 in. = 25.4 mm.

Volume changes due to moisture fluctuations should
be considered in design, especially for joint size.
Moisture permeability of stone veneers is generally not
a problem (Table 3.5.2). However, as stone veneers become thinner, water may penetrate in greater amounts

3

and at faster rates than normally expected, and damp
appearing areas of moisture on the exterior surface of
thin stone veneers will frequently occur. These damp
areas result when the rate of evaporation of water
from the stone surface is slower than the rate at which
the water moves to the surface.

3.5.12.3 Stone sizes
Stone veneers used for precast concrete facing are
usually thinner than those used for conventionally set
stone, with the maximum size generally determined by
the stone strength. Table 3.5.3 summarizes typical dimensions. Veneers thinner than those listed can result
in anchors being reflected on the exposed surface, excessive breakage, or permeability problems.
The length and width of veneer materials should be
sized to a tolerance of ±1/16 in. (±2 mm). This tolerance
becomes important when trying to line up the false
joints on one panel with those on the panel above or
below, particularly when there are a large number of
pieces of stone on each panel. Tolerance allowance for
out-of-square is ±1/16 in. (±2 mm) difference in length
of the two diagonal measurements.
Flatness tolerances for finished surfaces depend on
the type of stone and finish. For example, the granite
industry’s flatness tolerances vary from 3/64 in. (1 mm) for
a polished surface to 3/16 in. (5 mm) for flame (thermal)
finish when measured with a 4 ft (1.2 m) straightedge.
Tolerances should be clearly specified in the contract
documents. Thickness variations are less important,
because concrete will provide a uniform back face except at corner butt joints. In such cases, the finished
edges should be within ±1/16 in. (±2 mm) of the specified thickness. However, large thickness variations may
lead to the stone being encased with concrete and thus
restrict the relative movement of the materials.

Table 3.5.3 Dimensional parameters of various stone materials.

Stone Type

Recommended
thickness,
In. (cm)

Length range,
ft (m)

Width range, Maximum area,
ft (m)
sq ft (m2)

Marble

1.25 (3)

3-5 (0.9-1.5)

2-5 (0.6-1.5)

20 (1.9)

Travertine*

1.25 (3)

2-5 (0.6-1.5)

1-4 (0.3-1.2)

16 (1.5)

Granite

1.25 (3)

3-7 (0.9-2.1)

1-5 (0.3-1.5)

30 (2.8)

Indiana limestone

2 (5)

4-5 (1.2-1.5)

2-4 (0.6-1.2)

15 (1.4)

* Surface voids filled front and back.

ARCHITECTURAL PRECAST CONCRETE

|

213

3

SURFACE AESTHETICS

3.5.12.4 Design considerations / 3.5.12.5 Anchorage of stone facing

3.5.12.4 Design considerations
Structural design, fabrication, handling, and erection
considerations for veneered precast concrete units are
similar to those for other precast concrete wall panels,
except that special consideration must be given to the
veneer material and its attachment to the concrete.
The physical properties of the stone facing material
must be compared with the properties of the concrete
backup.
These properties include:
1. Tensile (axial and flexural), compressive, and shear
strength
2. Modulus of elasticity (axial tension, flexure, and
axial compression)
3. Coefficient of thermal expansion (Table 3.5.4)
4. Volume change
Because of the difference in material properties between natural stone and concrete, veneered panels are
more susceptible to bowing than homogeneous concrete units. Also, the flat surfaces of cut stone reveal
Table 3.5.4 Coefficients of linear thermal expansion of
aggregate and concrete.

Type of Rock
(Aggregate)

Average Coefficient
of Thermal Expansion
× 10-6/in./°F
Aggregate Concrete*

Quartzite, cherts

6.1–7.0

6.6–7.1

Sandstones

5.6–6.7

5.6–6.5

Quartz sands and gravels

5.5–7.1

6.0–8.7

Granites and gneisses

3.2–5.3

3.8–5.3

Syenites, diorites, and
andesite, gabbros, diabase,
and basalt

3.0–4.5

4.4–5.3

Limestones

2.0–3.6

3.4–5.1

Marbles

2.2–3.9

2.3

Dolomites

3.9–5.5

—

Expanded shale,
clay and slate

—

3.6–4.3

Expanded slag

—

3.9–6.2

Blast-furnace slag

—

5.1–5.9

* C oefficients for concretes made with aggregates from different sources vary from these values, especially those for
gravels, granites, and limestones. Fine aggregates generally
are the same material as coarse aggregates.

214

|

ARCHITECTURAL PRECAST CONCRETE

bowing more prominently than homogeneous concrete panels. However, there are a number of design
and production procedures to help minimize bowing.
For example, after the panels are erected, midpoint tieback connections can take out some bowing.

3.5.12.5 Anchorage of stone facing
The architect, engineer of record, and stone fabricator should conduct tests to determine anchor type
and spacing. This will allow the architect to provide
anchor spacing prior to bid so that common information can be supplied to all bidders (refer to ASTM
C 1242). The stone fabricator should drill the anchor
holes in the stone according to architectural specifications. Contract documents should clearly define type,
number, and location of anchors, and who supplies the
anchors.
A bondbreaker should be used between the stone
veneer and concrete backup in order to minimize bowing, cracking, and staining of the veneer. Connections
of natural stone to the concrete should be made with
flexible mechanical anchors that can accommodate
some relative in-plane movement.
Two methods may be used to prevent bond between
the veneer and concrete to allow for independent
movement:
1. A 6 to 10 mil polyethylene sheet.
2. A closed cell 1/8 to 1/4 in. (3 to 6 mm) polyethylene
foam pad. During shipment, consideration must
be given to preventing cracking of the stone due
to compressibility of the pad.
Preformed anchors, with a 5/32 in. (4 mm) minimum
diameter, fabricated from Type 304 stainless steel, are
supplied by the stone fabricator or, in some cases, by
the precaster depending on the contract document
requirements. The number and location of anchors
should be predetermined by a minimum of five shear
and tension tests conducted on a single anchor embedded in a stone/precast concrete test sample using
ASTM E 488 or ASTM C 1354 and the anticipated applied loads, both normal and transverse to the panel.
Loads anticipated during handling and shipping should
be included. Anchor size and spacing in veneers of
questionable strengths or with natural planes of weakness may require special analysis.
The number and location of anchor and size of the
stone should be based on specific test values for the

SURFACE AESTHETICS

3.5.12.5 Anchorage of stone facing

actual stone to be installed. Test samples for anchor
tests should be a typical panel section of about 1 ft2
(0.09 m2) and approximate as closely as possible actual
panel anchoring conditions. A bondbreaker should be
placed between stone and concrete during sample
manufacture to eliminate any bond between veneer
and concrete surface. Each test sample should contain
one anchor connecting stone to concrete backup and a
minimum of five tests are needed to determine tensile
(pull-out) and shear strength of each type of anchor.
Depending on the size of the project, it may be desirable to perform shear and tensile tests of the anchors
at intervals during the fabrication period.

3

Fig. 3.5.81 Typical anchor for marble veneer.
4"

2"

Depth of concrete
varies according to
design

db = 5/32" min.

2"
S.S. Anchor
30 to 45º

Bond
breaker

t varies
1-1/4" min. Preferred

6 to 9"
To edge of stone

Face of
stone veneer

1/2 t
3/4" min.

Four anchors are usually used per stone piece, with
Note: t = thickness of veneer; db = diameter of anchor
a minimum of two recommended. The number of
anchors has varied from one per 11/2 ft2 (0.14 m2) of
Fig. 3.5.82 Typical anchor for granite veneer.
stone to 1 per 6 ft2 (0.56 m2) with one per 2 to 3 ft2
(0.19 to 0.28 m2) the most common spacing. Anchors
should be 6 to 12 in. (150 to 300 mm) from an edge
S.S. Anchor
Depth of
with not more than 24 to 30 in. (600 to 750 mm) beconcrete
tween anchors depending on the local building code.
varies according
db = 5/32"
Bond
to design
min.
The shear capacity of the spring clip (hairpin) anchors
breaker
perpendicular to the anchor legs is greater than when
they are parallel (Table 3.5.5) and capacity depends on
1/2 t
t varies
the strength of the stone. A typical marble veneer an3/4" min.
Hole
1-1/4" min. Preferred
chor detail with a toe-in spring clip anchor is shown in
1/16" > db
30 to 45º
Fig. 3.5.81, while a typical granite veneer anchor detail
Face of stone veneer
is shown in Fig. 3.5.82. The toe-out anchor in granite
6 to 9"
may have as much as 50% more tensile capacity than a
To edge of stone
toe-in anchor, depending on the
stone strength.
Table 3.5.5 Ultimate shear capacity of spring clip (hairpin) anchors in granite from various sources*.
Depth of anchor holes should be
Shear parallel to anchor,
approximately half the thickness
lb (kg)
of the veneer (minimum depth
of 3/4 in. [19 mm]). Minimum
Stone
stone cover over the drilled hole
should be 3/8 in. (9 mm) to avoid
spalling during drilling and spotting from absorbed moisture. The
holes should be drilled at an an1
2400 to 2650 (1090 to 1200)
gle of 30 to 45° to the plane of
2
1800 (815)
the stone. Holes, approximately
3
1500 (680)
50% oversize, have been used to
4
2500
(1135)
allow for differential movement
5
2800 (1270)
between the stone and the con1
crete. However, holes /16 in. (2
6
3400 (1540)
mm) larger than the anchor are
7
1000 (455)
common, as excessive looseness
reduces holding power. Anchor *Need to apply safety factor.

Shear perpendicular to anchor,
lb (kg)

3200 to 3500 (1450 to 1590)
2500 (1135)
1500 (680)
3400 (1540)
4000 (1815)
4200 (1905)
1660 (725)

ARCHITECTURAL PRECAST CONCRETE

|

215

3

SURFACE AESTHETICS
3.5.12.5 Anchorage of stone facing

Fig. 3.5.83 Typical cross anchor dowels for stone veneer.

Depth of
concrete
varies
according
to design

t (Varies)

S.S. Anchor
db = 3/16 to 5/8"
2-1/2" Min.
db

2/3 t
2" Max

45 to 60º

Bond
breaker

Close slots
with tape

2"

Hole
1/16" > db

Face of
stone veneer

Dowels may
be epoxied

Note: t = thickness of veneer; db = diameter of anchor

holes should be within ±3/16 in. (±5 mm) of the specified
hole spacing, particularly for the spring clip anchors.
Stainless steel dowels, smooth or threaded, may be
installed to a depth of 2/3 of the stone thickness, with a
maximum depth of 2 in. (50 mm) at 45 to 60° angles
to the plane of the stone. The minimum embedment
in the concrete backup to develop the required bond
length is shown in Fig. 3.5.83. Dowel size varies from
3
/16 to 5/8 in. (5 to 16 mm) for most stones, except that
it varies from 1/4 to 5/8 in. (6 to 16 mm) for soft limestone and sandstone; it depends on the thickness and
strength of the stone.

Fig. 3.5.84 Typical anchors for limestone veneer.
Deformed
S.S. Anchor db = 3/16 to 5/8"
Concrete
Dowels may
be epoxied

Varies

In some climates two-part polyester or epoxy is placed
in the anchor holes in order to eliminate moisture condensation in the holes and the possible dark, damp
appearance of moisture on the exposed stone surface.
The polyester or epoxy increases the shear capacity
and rigidity of the anchors. The rigidity may be partially overcome by using 1/2 in. long (13 mm) compressible (60 durometer) rubber or elastomeric grommets or
sleeves on the anchor at the back surface of the stone
(Fig. 3.5.85).

Varies
3/4" min.
in 2" stone;
2" max.
in 5" stone

Deformed
S.S. Anchor db = 3/16 to 5/8"

15º

Bond breaker

Concrete

Varies

2 to 5"

Stone
1 to 2"
*Use anchors at opposing angles

|

Some flexibility should be introduced with all anchors by minimizing the anchor’s diameter to allow for the inevitable relative movements that occur
with temperature variations and concrete shrinkage.
Unaccommodated relative movements can result in
excessive stress problems and eventual failure at an
anchor location. Consideration may be given to accelerated cyclic temperature tests on the stone-concrete
assembly to determine the affect of strength loss on
the shear and tensile strengths of the anchors.

Bond breaker

Hole 1/16" > db
Stone

216

Limestone also has been used traditionally with a
thickness of 3 in. (75 mm), but it is now being used as
thin as 13/4 in. (44 mm), although one limestone group
recommends a minimum of 2 in. (50 mm). With limestone, use a bondbreaker, along with mechanical anchors. Dowels and spring clip anchors can be used to
anchor limestone. Typical dowel details for limestone
veneers are shown in Fig. 3.5.84; the dowels should
be inserted at opposing angles to secure stone facing
to backup concrete.

30 to 45º

ARCHITECTURAL PRECAST CONCRETE

Fig. 3.5.85 Example of a compressible sleeve used to reduce
stone anchor rigidity when the anchors are epoxied in the
stone.

SURFACE AESTHETICS

3.5.12.5 Anchorage of stone facing / 3.5.12.7 Veneer jointing

It may be more desirable to fill the anchor hole with
a low modulus polyurethane sealant. The overall effect
of either polyester, epoxy, or sealant materials on the
behavior of the entire veneer should be evaluated prior
to their use. At best, the long-term service life of adhesive-embedded anchors is questionable, so any increase
in pull-out strength of the anchors should not be used
in calculating long-term anchor capacity. When using
polyester or epoxy in anchor holes, the precaster needs
to follow the manufacturer’s recommendations as to
mixing and curing temperature limitations.
The stone trade associations and the suppliers of the
various kinds of building stones recommend safety
factors. Due to the variation in the physical properties of natural stones and to account for the risks of
brittle failure and the effects of weathering, there are
more recommended safety factors than those used
for manufactured building materials, such as steel
and concrete. The minimum recommended safety factor, based on the average of the test results, is 4 for
anchorage components in stone. If the range of test
values exceeds the average by more than ±20%, then
the safety factor should be applied to the lower bound
value (see the Appendix to ASTM C 1242 for a discussion on safety factors).

3.5.12.6 P
 anel watertightness
The bondbreaker between the stone veneer and concrete backup may function as a vapor barrier on the
concrete’s exterior face, keeping moisture in the veneer or at the interface unless drainage provisions are
provided. After some period of time, gaps also may develop between the stone veneer and concrete backup
at the bondbreaker. These gaps could allow moisture
penetration due to capillary action and gravity, particularly where the window or roof design allows water to

Fig. 3.5.86 Effect of changes in the sand aggregate binder ratio
on the thermal coefficient of an epoxy.

Thermal coefficient x 10-6 in/in/deg F

Differential thermal expansion of the stone and unfilled epoxy (without sand) may cause cracking of the
stone veneer. Epoxies yield under stress and, if properly formulated, will accommodate relatively large
dimensional changes resulting from thermal effects.
The coefficient of expansion of the stone and epoxy
should closely match. However, this may be overcome
by keeping the oversizing of the hole to a minimum,
thereby reducing epoxy volume and using stone flour
or fines, or fine sand as a filler for the epoxy to reduce
the coefficient of thermal expansion of the epoxy and
the shrinkage (Fig. 3.5.86).

3

48
42
36
30
24
18
12
6

Sand-filled epoxy
Concrete

0
1

2

3

4

5

6

7

8

9

10

Aggregate-binder ratio

flow over the top of the panel. One solution to overcome this problem is a two-stage joint. This approach
provides an airtight 1 in. (25 mm) wide urethane seal,
bonded to the stone veneer and concrete backup, and
continuous along both sides and top of the panel.
Other designers have used a sealant applied to the top
and side edges of the stone/concrete interface after
the panels are cast. Care must be taken to ensure that
the sealant used is compatible with the sealant to be
applied to panel joints after erection of the panels. The
bondbreaker should not be sealed at the bottom of
the panel. This ensures any moisture that penetrates
behind the stone veneer can drain freely.

3.5.12.7 Veneer jointing
In the form, the stone veneer pieces are temporarily
spaced with a non-staining, compressible spacing material, such as rubber, neoprene, or soft plastic wedges,
or a chemically neutral, resilient, non-removable gasket, such as sealant backer rod, which will not stain
the veneer or adversely affect the sealant to be applied
later. Shore A hardness of the gasket should be less
than 20.
The gaskets should be of an adequate size and configuration to provide a recess to receive the sealant and
also prevent any of the concrete backup from entering
the joints between the veneer units. Non-acidic based
masking or duct tape (other types will stain stone) may
also be used to keep concrete out of the stone joints so
as to avoid limiting stone movement. Spacer material
should be removed after the panel has been removed
from the mold unless it is a resilient sealant backup.

ARCHITECTURAL PRECAST CONCRETE

|

217

3

SURFACE AESTHETICS

3.5.12.7 Veneer jointing / 3.5.12.9 Applications

Joints between veneer pieces on a precast concrete
element are typically a minimum of 3/8 in. (10 mm),
although they have often been specified equal to
the joint width between precast concrete elements.
Because actual joint width between precast concrete
panels (as erected) depends largely on the accuracy of
the main supporting structure, it is not realistic to require matching joint widths between stone pieces and
between panels.
The use of an invisible joint (for example, less than
/8 in. [10 mm]) is not recommended because the joint
must have the width necessary for sealant design to allow for movements, tolerances, and other dimensional
or volumetric changes. Also, due to tolerances and natural warping, adjacent panels may not be completely
flush at the joint, and shadow lines will appear. Rather
than attempting to hide the joint, the joint should be
emphasized by finding an aesthetically pleasing joint
pattern with a complementary joint size.

3

When stone veneer is used as an accent or feature
strip on precast concrete panels, a 1/2 in. (13 mm) space
is left between the edge of the stone and the precast
concrete to allow for differential movements of the
materials. This space is then caulked as if it were a conventional joint.
The sealant between stones or panels should be an
elastomeric, usually urethane, polysulfide, or silicone,
which will not stain the stone-veneer material. Some

grades of silicone sealants are not recommended by
their manufacturers for applications on high-calcite content stone, as they may stain light-colored stones or may
cause a change in surface moisture–absorption characteristics that can be seen whenever the stone is wet.

3.5.12.8 Repair
Should minor damage occur to the veneer stone during shipping, handling, or erection, field remedial work
can be performed successfully. The precaster normally
does such repairs, with repair procedures developed in
consultation with the stone fabricator. If it is necessary
to replace a stone piece, satisfactory techniques have
been developed when the back of the panel is accessible or after the panels have been erected and the
back of the panel is inaccessible.

3.5.12.9 Applications
Over the last 40 years, many structures have been
constructed with stone veneer–faced precast concrete
panels. Several examples are shown to illustrate the
use of the various types of stone.
Marble: The base structure of the temple in Fig.
3.5.87(a) consists of 11/4 in. (3 cm) Vermont marble
facing backed with 4 in. (100 mm) of precast concrete. Various stone clad panel shapes are shown in
Fig. 3.5.87(b) and (c).

Fig. 3.5.87(a), (b) & (c)
Portland, Oregon Temple for the Church of Jesus Christ
of Latter Day Saints, Lake Oswego, Oregon;
Architect: Lee, Ruff, Stark Architects;
and Leland Gray Architects.

(b)
(c)

(a)

218

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.12.9 Applications

3

(a)

(b)
Fig. 3.5.88(a) & (b)
Joslyn Art Museum
Omaha, Nebraska;
Architect: Sir Norman Foster and Partners, Design Architect;
Henningson, Durham & Richardson Inc., Architect of Record;
Photos: (a) HDR Architecture, (b) Patrick Drickey.

Architectural precast concrete panels
for the new wing of the museum in Fig.
3.5.88(a) are clad in pink 11/4 in. (3 cm)
Etowah Fleuri Georgia marble to match
the original stone building constructed in
1931. Labor and material costs were reduced using this system compared to traditional stone cladding systems. There were
199 panels, with the heaviest piece weighing 22,100 lb (10,000 kg) (Fig. 3.5.88[b]).
The precast concrete panels were stacked
and laterally tied to the lightweight, structural steel structure.
Travertine: Over 73,000 pieces of 11/4
in. (3 cm) buff travertine were anchored to
7055 precast concrete panels to produce
600,000 ft2 (55,800 m2) of cladding for a
37-story administrative tower (Fig. 3.5.89)
and a 14-story comptrollers building. The
precast concrete units—wall panels, spandrels, and column covers—measure up to
5 ft (1.5 m) wide and 20 ft (6.1 m) long
and weigh up to 6000 lb (2700 kg) each.
Fig. 3.5.89
SBC Headquarters
Dallas, Texas;
Architect: JPJ Architects;
Photos: (main) Wes Thompson Photography,
(inset) REES.

ARCHITECTURAL PRECAST CONCRETE

|

219

3

SURFACE AESTHETICS
3.5.12.9 Applications

Fig. 3.5.90
Roseville Telephone Company
Roseville, California;
Architect: Williams + Paddon;
Photo: Ed Asmus.

Sandstone: Architectural precast concrete panels
were integrally cast with 11/4 to 13/4 in. (3 cm to 44
mm) clefted Arizona Red Sandstone on 4-in.-thick
(100 mm) concrete backing for the two-story structure,
which added 78,000 ft2 (7200 m2) to an existing fourbuilding campus (Fig. 3.5.90). Unique features include
step-back chamfer detailing at corners and bullnoses.
Window recesses and light shelves provide solar control and natural daylighting.

220

|

ARCHITECTURAL PRECAST CONCRETE

The lower three floors of the office building in Fig.
3.5.91(a) has 2 in. (50 mm) of red sandstone cast integrally in the 4-in.-thick (100 mm) precast concrete
column covers and spandrels, which also have lightly
sandblasted concrete areas (Fig. 3.5.91[b]).
Granite: The 26-story flatiron building rising from an
18,000 ft2 (1700 m2) triangular site is an outstanding
example of stone veneer–faced precast concrete work.

SURFACE AESTHETICS

3.5.12.9 Applications

3

(a)

(b)

Fig. 3.5.91(a) & (b)
Collier Center, Phoenix, Arizona; Architect: Opus Architects & Engineers.

ARCHITECTURAL PRECAST CONCRETE

|

221

3

SURFACE AESTHETICS
3.5.12.9 Applications

(b)

Fig. 3.5.92(a) & (b)
388 Market Street
San Francisco, California;
Architect: Skidmore Owings & Merrill;
Photos: Skidmore Owings & Merrill NYC.

(a)

222

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.12.9 Applications

The mixed use building is clad with 1915 precast concrete panels of varying size that are faced with 11/4 in. (3
cm) thick Imperial red granite (Fig. 3.5.92[a] and [b]).
The corporate office high-rise building in Fig. 3.5.93
features 1400 architectural precast concrete panels

3

with embedded polished granite to achieve a level of
detail that was not practical using handset granite.
Acid-etched precast concrete resembling limestone
was used at exposed corners on the street level, and
throughout the tower where more intricate detail was
desired.

Fig. 3.5.93
Campanile, Atlanta, Georgia; Architect: Thompson, Ventulett, Stainback & Associates Inc.; Photo: E. Allan McGee Photography.

ARCHITECTURAL PRECAST CONCRETE

|

223

3

SURFACE AESTHETICS
3.5.12.9 Applications

Fig. 3.5.94
Terry Sanford Institute of Public Policy,
(Duke University West Campus),
Durham, North Carolina;
Architect: Architectural Resources Cambridge Inc.;
Photo: Jonathan Hillyer.

Limestone: Many college campuses feature a general
architectural theme and many variations on that theme.
Due to changing tastes, cost constraints, and material availability, a variety of styles are created through
the years. The building in Fig. 3.5.94 re-established a
strong design sense. The older campus buildings are
richly detailed with pitched roofs, gothic towers, and
window tracery. The building in Fig. 3.5.94 manages
to recall the original collegiate gothic architecture of
the campus while simplifying the construction process.
Approximately 4500 pieces of gray, mottled, German
(Jura) limestone were cast into the architectural precast
concrete panels during fabrication and were anchored
into the panels by stainless steel pins. The exposed surfaces used in detailing the gables, towers, and window
framing were given a light sandblast finish to resemble

224

|

ARCHITECTURAL PRECAST CONCRETE

sandstone. The German limestone replicates the coloring and texture of the original campus stone, connecting itself both visually and symbolically to the old
campus.
The building in Fig. 3.5.95 is clad with 13/4 in. (44 mm)
thick polychrome limestone supported on 2208 precast
concrete panels. The beige and white-hued limestone
fits comfortably within downtown, which is renowned
for its flamboyant historical architecture. Precast concrete was selected as the backing for the limestone
over other systems because of the plastic shaping possibilities that allowed substantial in-and-out relief in
the exterior plane. Windows, for instance, are set back
from the face of the building by an average of 16 in.
(400 mm). The cost to achieve this degree of modeling

SURFACE AESTHETICS

3.5.12.9 Applications

3

Fig. 3.5.95
GSA Federal Building,
Oakland, California;
Architect: Kaplan McLaughlin Diaz;
Photo: Kaplan McLaughlin Diaz.

in the exterior skin was significantly less with precast
concrete panels as opposed to other systems.

(b)

The project in Fig. 3.5.96(a) is comprised of limestone
slabs cast on precast concrete panels. Precast concrete
also forms the horizontal sunshade elements that provide shading relief to the interior perimeter spaces. The
precast concrete panels became the structural system
that carried the limestone between structural columns.
The lower portion of the building (Fig. 3.5.96[b]), clad in
(a)

Fig. 3.5.96(a) & (b)
University of Chicago Graduate School of Business, Chicago, Illinois;
Architect: Rafael Vinoly Architects, P.C.; Photos: Brad Feinknopf,
Rafael Vinoly Architects.

ARCHITECTURAL PRECAST CONCRETE

|

225

3

SURFACE AESTHETICS
3.5.12.9 Applications

(a)

a horizontal pattern of limestone panels, establishes
the scale of the base of the building and echoes the
horizontal composition of the neighboring Robie
House.
Quartzite: The chapel portion of the hospital in
Fig. 3.5.97(a) has concave precast concrete panels
with 1 in. (25 mm) thick polished quartzite inlays cast
into the panels. This saved thousands of dollars by
not having to erect thousands of little pieces; it also
reduced the material thickness. The inset thin brick
panels are trimmed with acid-etched limestone finish
to match a cut stone appearance. The brick panels
on the main wall of the north hospital elevations are
made with a convex radius to give a more distinct
look and to set it off from the chapel. The medical
office is comprised of both the brick and acid-etched
panels on the upper levels and quartzite panels at the
base and forming the columns, thereby blending into
the other two parts of the project (Fig. 3.5.97[b]).

226

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 3.5.97(a) & (b)
St. Clare’s Hospital, Weston, Wisconsin;
Architect: Hammel, Green & Abrahamson Inc.;
Photos: Staley Studio.

(b)

SURFACE AESTHETICS

3.5.12.9 Applications

3

Fig. 3.5.98
Liberty Place – Phase II, Philadelphia, Pennsylvania; Architect: Zeidler Roberts Partnership Inc.; Photo: Zeidler Roberts Partnership Inc.

Accents or Feature Strips: There are a variety of
ways that stone veneer can be used as an accent or feature strip on precast concrete panels. Two approaches
to accent or feature strip applications are shown in the
following projects.
The final phase of the mixed-use complex in Fig.
3.5.98 encompasses a retail podium and a hotel. The
first phase office tower was clad in bands of blue-pearl

granite with silver and blue reflective glass, setting the
context for the second phase. Granite bands and accents are employed in combination with sandblast-finished precast concrete panels. This resulted in an extremely viable alternative to the much more expensive
option of exclusively using granite. The pattern of large
reveals corresponds to the bands of the first phase and
facilitated the installation of the granite bands and
accents.

ARCHITECTURAL PRECAST CONCRETE

|

227

3

SURFACE AESTHETICS
3.5.12.9 Applications

Some fourteen-hundred 6-in.-thick (150 mm) architectural precast concrete panels clad the hospital facility in Fig. 3.5.99(a). Red Spanish granite insets, 11/4
in. (3 cm) thick, with one anchor per 2 ft2 (0.19 m2),
were provided to create an intermittent horizontal dark
band running the length of the building at every story (Fig. 3.5.99[b]). This reflects the surrounding brick
architecture of the neighborhoods. Green terra cotta
medallions, interspersed among the granite bands at
spandrel intersections, protrude out from the wall to
add further detail and interest. Two different finishes
were provided through different levels of sandblasting to produce contrasting finish textures around the
windows, with reveals added between and around the
insets and bands. Granite inserts highlight the light
acid-etched panels in Fig. 3.5.100.
Fig. 3.5.99(a) & (b)
Shriners Hospital for Crippled Children,
Sacramento, California;
Architect: Odell Associates and HDR, Associate Architect.

(b)

(a)

228

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.13 Applied Coatings

3

Fig. 3.5.100
Walsh Library at Seton Hall University, South Orange, New Jersey (1994); Architect: Skidmore Owings & Merrill; Photo: Eduard Hueber/
Archphoto.com.

3.5.13 A
 pplied Coatings
Because architectural precast concrete is a highstrength, durable product with integral color, either
from the aggregates, cement, or pigments, it does not
require staining or painting. Integrally colored concrete
can provide significant long-term savings over the cost
of surface coatings because the ongoing maintenance
expense is eliminated. However, some projects use
staining or painting for various reasons. For example,
pigmented stains may be used to color the surface and
blend in any panel discoloration to produce a more uniform appearance. If the entire panel is to be coated,
it may be acceptable to permit a gray concrete that
would not necessitate color control (basically a lesser

quality of finish). In other words, the quality of finish
color would not be expected to be a high-quality architectural finish.
Every paint and stain is formulated to provide certain
performance characteristics under specified conditions.
Because there is a vast difference in paint or stain types,
brands, prices, and performances, knowledge of composition and performance standards is necessary for obtaining a satisfactory paint or stain for use on concrete.
The quality of paint for concrete is not solely determined by the merits of any one raw material used in
its manufacture. Many low-cost paints with marginal
durability are on the market. In order to select proper

ARCHITECTURAL PRECAST CONCRETE

|

229

3

SURFACE AESTHETICS
3.5.13 Applied Coatings

Fig. 3.5.101
Lake County Office Center, Deerfield, Illinois;
Architect: Skidmore, Owings & Merrill;
Photo: Holmstrom Photography.

paints, the architect should consult with manufacturers that supply products that are known for their
durability and, if possible, obtain technical data from
them that explain the chemical composition and types
of coatings suitable for the specific job at hand. For
high-performance coatings, proprietary brand-name
specifications are recommended.
Paints are a mixture of pigment, which hides the surface, and resin, which binds the pigment together. The
amount of pigment in proportion to the resin and the
type of resin will affect the fluidity, gloss, and durability
of the paint.
The pigment volume concentration (PVC) compares
the amount of pigment in a paint to the amount of
binder. As the PVC increases, the paint has more pigment and less binder. High PVC paints (45 to 75%)
generally brush on easier, have greater hiding power,
and usually cost less than low PVC paints. Low PVC
paints (10 to 22%) are generally more flexible, durable, and are glossier.

230

|

ARCHITECTURAL PRECAST CONCRETE

Coatings applied to exterior surfaces should be of
the breathing type, permeable to water vapor but impermeable to liquid water. Typically, latex paints are
suitable for most exterior applications. Typically, latex
paints or epoxy, polyester, or polyurethane coatings
may be applied to the interior surface of exterior walls
if a vapor barrier (paint or other material) is necessary.
See Section 3.5.16 for finishing procedures for interior
surfaces to be painted.
The coating manufacturer’s instructions regarding
mixing, thinning, tinting, surface preparation, and application should be strictly followed.
The smooth gray finish on the spandrel panels in Fig.
3.5.101 was painted at the jobsite in tan and maroon
and the ends of the large bullnoses were gilded with
gold leaf.
The designers of the dormitory (Fig. 3.5.102[a] & [b])
sought to visually break up the building’s massive appearance by using step-backs and different colors. On

SURFACE AESTHETICS

3.5.13 Applied Coatings / 3.5.14 Architectural Trim Units

the first two floors, a band of color breaks up the sevenstory height-a two-story band with three individual colors on each bay as a step-back, then a band across the
top.

3

(b)

Architectural precast concrete panels with a brick
form liner, stained in the plant to control quality, were
selected to simplify construction, improve cost efficiency, and shorten construction time (Fig. 3.5.103). The
exterior surface treatment of each panel consists of
fields of brick pattern subdivided by horizontal bands
of what appears to be stone projecting 1 in. (25 mm).
At the base of each 10-ft-wide (3 m) panel, an additional form liner type is used to give the appearance of
a rusticated base.

3.5.14 A
 rchitectural Trim Units
Cast stone is a type of architectural trim (AT) unit,
manufactured to simulate natural cut stone, used in
masonry applications. It is generally used as ornamentation and architectural trim for stone bands, sills, lin(a)
Fig. 3.5.102(a) & (b) Residence Hall One,
University of South Florida,
St. Petersburg, Florida;
Architect: KBJ Architects Inc.;
Photos: Clay Callahan.

ARCHITECTURAL PRECAST CONCRETE

|

231

3

SURFACE AESTHETICS
3.5.14 Architectural Trim Units

Fig. 3.5.103
Lock-Up Self Storage Center,
Park Ridge, Illinois; Architect: Sullivan Goulette Wilson, Ltd. formerly Sullivan Goulette, Ltd.; Photo: Sullivan Goulette Wilson, Ltd.

tels, copings, ashlar veneer, balustrades, and door and
window surrounds, where natural cut stone may otherwise be used in masonry wall systems.
Because it is a “building stone,” architectural trim units
are usually specified under the masonry section (047200)
rather than section 034500 of the project specifications
where architectural precast concrete is normally specified. Architectural trim units are usually installed by the
mason contractor, rather than the precast concrete erector. Because of different setting procedures used for
each application, projects utilizing both materials should
have specifications governing each.
The two most widely used casting methods today are
the dry tamp and wet cast methods. Both require a
carefully proportioned mixture consisting of graded and
washed natural sand or manufactured sands of granite, marble, quartz, or limestone meeting the requirements of ASTM C 33, except that gradation may vary
to achieve the desired finish. A 1:3 cement-aggregate
ratio mixture is proportioned for a maximum density

232

|

ARCHITECTURAL PRECAST CONCRETE

and to meet the surface finish requirements. Units may
be either homogenous throughout or consist of a face
mixture and a backup mixture. A fine-grained texture
similar to natural cut stone, with no bug/blow holes or
air voids in the finished surface, is produced.
In the dry tamp method, a pneumatic machine rams
and vibrates moist, zero-slump concrete against rigid
formwork. When the concrete is densely compacted,
it is removed from the form and cured overnight in
a moist, warm room. The limitation of this method is
that it generally requires one flat, unexposed side in
the design of the basic section. L-shaped, U-shaped, or
cored-out stones can be made; however, the designer
should be aware that these specially shaped units must
be hand-molded and will result in higher unit costs.
Where return legs are essential to the design, the
depth of returns should be standardized at 6 or 8 in.
(150 or 200 mm) for ease of manufacture. So long as
the return leg equals the depth of the section, no additional cost is incurred (Fig. 3.5.104).

SURFACE AESTHETICS

3.5.14 Architectural Trim Units

3

Location of dowel holes, anchor slots, flashing
grooves, false joints, and similar features

Fig. 3.5.104 Returns in cast stone.

Formed sides of unit. . . . . . . . . . . . . .  1/8 in. (3 mm)
DON’T

DO

OR

OR

The wet cast method is similar to the production process for architectural precast concrete. Mixture designs
usually have a maximum 3/8 in. (10 mm) coarse aggregate and are comprised of an abundance of fines, typically 15% very fine sand. L-shaped stones should be
avoided as it is difficult to produce a good finish with
no bugholes on these shapes.
To help strip the stone from the mold, a minimum 1:8
draft should be designed for all surfaces perpendicular
to the face pattern. This minimizes the number of mold
pieces required to create the face pattern. Negative
draft profiles should be avoided whenever possible.
Cast stone should have a minimum thickness of 21/2
in. (63 mm) to reduce stripping, handling, and packaging costs. A 21/2 in. stone generally costs the same as a
4-in.-thick (100 mm) stone. The thickness of a projection should be twice the length of the projection.
Dimensional tolerances are as follows:
Cross-section dimensions . . . ±1/8 in. (±3 mm)
Length (in inches). . . . . . . . . L /360 or ±1/8 in. (±3 mm)
whichever is greater, not
to exceed ±1/4 in. (±6 mm)
Maximum length . . . . . . . . . <15 times effective
cross-section thickness
Warp, bow, or twist. . . . . . . . . L/360 or ±1/8 in. (±3 mm),
whichever is greater

Unformed sides of unit. . . . . . . . . . . .  3/8 in. (9 mm)
The compressive strength of cast stone should be a
minimum of 6500 psi (45 MPa), determined by testing
2 in. (50 mm) cubes in accordance with ASTM C 1194
(equivalent to about 5000 psi (34.5 MPa) for 6 in. [150
mm] cylinders). The units should have a cold water
absorption of less than 6% by weight when tested
in accordance with ASTM C 1195, Method A, or less
than 10% when tested in accordance with Method B.
The amount of reinforcement in units with a dimension greater than 12 in. (300 mm) in more than one
direction should not be less than 0.25% of the crosssectional area.
Architectural trim units produced by the wet cast
method allow the use of air entrainment to obtain freezing and thawing durability. The air content for wet cast
products used in a freezing and thawing environment
should be 4 to 6%. For dry tamp products, freezing and
thawing resistance may be tested according to ASTM
C 666, Procedure A, as modified by ASTM C 1364
(weight loss 5% after 300 cycles). Also, sills, copings, and
projecting courses should have a slope (wash surface) of
1:12 (25 mm in 300 mm) to allow water runoff on exposed top surfaces to prevent saturation of the unit.
Cast stone is available in a variety of colors, shapes,
and finishes. Typical colors will match either natural
stone, brick, or terra cotta. The support system and
setting techniques will influence the size of the cast
stone pieces. Cast stone is normally anchored to masonry or concrete walls or a steel stud grid. The size
limitations of cast stone are about the same as those
of natural cut stone, usually about 3 to 6 ft (0.9 to 1.8
m) long. Length should be kept within 15 times the
maximum cross-section thickness whenever possible.
Longer lengths invite cracking and handling problems.
Similar to architectural precast concrete, the key to optimum economy is repetition of ornament. Cast stone
is typically given an acid-etch finish and the surface texture matches cut stone (sand texture, cleft or stippled
face, cut tooled, or bushhammered). In some cases,
the stone is given a rubbed or abrasive blasted finish to
expose the aggregates and simulate limestone.
When cast stone is used as part of a brick veneer system, consideration must be given to the differential

ARCHITECTURAL PRECAST CONCRETE

|

233

3

SURFACE AESTHETICS
3.5.14 Architectural Trim Units

temperature and moisture volume movements of the
various materials by limiting lengths to prevent mortar
debonding and by proper location of control joints. A
bond break should be provided between clay brick and
cast stone banding to accommodate the differential
movement that will occur. Flashing is often placed either directly above or below the banding course. When
the accent band consists of more than one course, a
crack in a head joint will frequently continue through
the body of the unit located above or below. Therefore,
a single course accent band is recommended or stack
bond should be designed if more than one course is
desired.
Curved sections should be kept shorter than 4 ft (1.2
m) whenever possible and the major unexposed back
surface should be flat. Sufficient clearance in the masonry wythe or structural wall section should be provided. A sill section generally presents no problem when
designed with a radius front and rear because the major
unexposed side is flat at the masonry bed joint.

Through-wall flashing and weep holes should be used
at all interruptions in the wall, such as at window heads
and relieving angles, or at a change in material, such
as stone to brick. Flashing consisting of stainless steel,
EPDM, or rubberized asphalt must be continuous and
properly lapped and sealed at the base of the wall and
at relieving angles. When flashing is used over openings, such as at windows, end dams are required. In the
case of a masonry backing wythe, the flashing should
be turned up a minimum of 8 in. (200 mm) and extend
into the masonry backing.

Not all joints between cast stones or between cast
stone and other materials should be filled with Type N
mortar (ASTM C 270). All head joints at coping stones
and joints at column covers, cornices, platforms, soffits,
window sills and in general, all stone sections with projecting profiles, exposed top joints, or rigid suspension
connections to the supporting structure should be “soft”
sealant joints. Mortar joints are best suited for masonrybound trim items such as belt courses, lintels, window
surrounds, date stones, inscription blocks, quoins, keystones, and similar applications. Anchors, ties, and flashing are built into mortar joints as units are set.

Open weep hole head joints of at least 1 in. (25 mm)
are recommended. They should be spaced no more than
24 in. (600 mm) apart. Rope wicks can also be used,
but weep holes should be placed closer together, at 16
in. (400 mm) on center, because wicks do not drain as
quickly. Plastic tubes are not recommended because
they are easily clogged by mortar or insects. In stones
over 24 in. (600 mm) in length, a 3/8-in.-wide (10 mm)
by 1-in.-high (25 mm) notch through the base of the
stone is recommended for drainage. Unnecessarily long
lengths of stone are discouraged because adequate
drainage between weep holes can be a problem.

Regardless of whether mortar or sealant is selected as
the face joint material, the mortar must be raked out of
the joint to a minimum depth of 3/4 in. (19 mm). Mortar
joints should be pointed by placing and compacting
mortar in layers not greater than 3/8 in. (10 mm). Each
layer should be compacted thoroughly and allowed to
become thumbprint-hard before applying next layer.
Exposed joints should be tooled slightly concave when
thumbprint-hard using a jointer larger than joint thickness. If sealant is to be used at the head joints, then
no mortar should be used there. Sealant joints allow
for movement at the vertical joints. Pointing is required
because mortar shrinks and cracks, which can cause
leaks.

Copings perform best when the mortar bond with the
masonry wall is maintained. For this reason, flashing
should not extend over the full width below the coping. Instead, the flashing should be turned down into
the drainage cavity and then out through the exterior
supporting wythe below. To minimize the chance of
water intrusion, bridge coping stones over vertical control joints (Fig. 3.5.105). Fill the joints between coping
stones with sealant and place sealant, instead of mortar,
in the bed joint from the control joint to the nearest
joint in the coping.

Bed (collar) joints in most hand-set stones may be set
with the usual wet consistency mortar used in setting

234

|

masonry. Stiffer mortar must be used when setting larger stones, and plastic or lead shims are recommended
for all pieces over 300 lb (136 kg). When setting, fill all
dowel holes, anchor slots, and similar building stone anchor pockets completely with mortar. Non-shrink grout
may be specified for dowel connections. Accent band
bed joints should be raked and caulked to encourage
cracks in the mortar joints rather than through the cast
stone units.

ARCHITECTURAL PRECAST CONCRETE

For optimum economy, standard building stone anchors should be used whenever possible. Two anchor
slots are provided in the top of most trim stones to receive anchor straps. Alternatively, a continuous anchor
slot may be used. This eliminates the need to locate

SURFACE AESTHETICS

3.5.14 Architectural Trim Units

3

ning bond of the brickwork and must match brick
coursing in height, minus one joint.

Fig. 3.5.105 Coping joints.
Fill joint and dowel holes with sealant

Sealant

anchors in the field. Dowel holes, continuous relieving
angles, and anchor straps are preferred over embedded inserts and weld plates. A continuous relieving angle is the most economical support system. The stone
anchors and other accessories are all furnished by the
masonry or general contractor or the stone setter. For
anchors, dowels, and ties, it is necessary to specify
the type of material, corrosion resistance (Eraydo alloy
zinc, galvanized steel, brass, or stainless steel, Type 302
or 304), and dimensions. The anchors may be required
to penetrate flashing to allow a secure connection to
the structure. Where this occurs, proper steps must be
taken to ensure a watertight connection at the interface so that the anchor does not compromise the integrity of the flashing. Grommets, thimbles, sleeves,
couplings, and sealants are available for this purpose.
Raised fillets (lugs) should be provided at backs of sills
and at ends indicated to be built into jambs.
Because cast stone is a masonry product that is usually built into a brickwork façade, brick coursing tables
must be used in determining the sizes of the units.
When determining the height of cast-stone window
heads and sills, the following should be considered:
1. T he bottom of the sill to the bottom of the lintel
must always equal a brick coursing dimension.
2. T he lintel height (if stone) equals brick coursing
dimension minus one joint.
3. T he sill height equals a brick coursing dimension
to the bottom of the lintel, minus the window
dimension.
4. T he height of the lug must equal a brick coursing
dimension minus one joint.
5. Q
 uoins should be sized in length to match the run-

Slip sills have no lugs and the lengths are figured 1/2
in. (13 mm) less than the brick masonry opening (for
a 1/4 in. [6 mm] mortar joint) or 3/4 in. (19 mm) less
where sealant is desired to provide a 3/8 in. (10 mm)
joint. Window sill joints are centered under mullions.
Drips should be provided on projecting elements, such
as heads and sills, to prevent staining.
When cast stone is used as part of a brick façade,
consideration must be given to the differential movement of the two materials by proper location of control
joints. All brick veneer control joints should extend into
the cast stone.
Extreme care should be used when cleaning a brick
façade to avoid acid runoff onto the architectural trim
units.
Installation tolerances are as follows:
1. Variation from plumb. . . . . . . 1/8 in. in 10 ft (3 mm
in 3 m) or 1/4 in. in
20 ft (6 mm in 6 m)
or more.
2. Variation from level. . . . . . . . . 1 /8 in. in 10 ft (3 mm
in 3 m), 1/4 in. in 20
ft (6 mm in 6 m),
or 3/8 in. (10 mm)
maximum.
3. Variation in joint width . . . . . . . 1 /8 in. in 36 in. (3 mm
in 900 mm) or 1/4
of the nominal joint
width, whichever is
less.
4. Variation in plane between
adjacent surfaces (Lipping) . . . . . . . 1/8 in. (3 mm).
Acceptance criteria are as follows:
1. All surfaces intended to be exposed to view should
have a fine-grained texture similar to natural stone,
with no air voids in excess of 1/32 in. (0.8 mm) and
the density of such voids shall be less than three
occurrences per any 1 in.2 (25 mm2) and not obvious under direct daylight illumination at a distance
of 5 ft (1.5 m). Visible cracks exceeding 0.005 in.
(0.13 mm) are not acceptable.
2. Units should exhibit a texture approximately equal
to the approved sample when viewed under direct

ARCHITECTURAL PRECAST CONCRETE

|

235

3

SURFACE AESTHETICS
3.5.14 Architectural Trim Units

daylight illumination at a 10 ft (3 m) distance. Also,
minor chipping should not be obvious under direct
daylight illumination at a distance of 20 ft (6 m).
Testing for color variation should be performed according to ASTM D 2244. The permissible variations in color between units of comparable age
subjected to similar weathering exposure are:
a. Total color difference – not greater than 6 units.
b. Total hue difference – not greater than 2 units.
Figures 3.5.106 and 3.5.107 show typical applications of cast stone. These include window surrounds
such as sills, mullions, jambs and headers, entry fea-

Fig. 3.5.106
Park Cities Baptist Church, Dallas, Texas;
Architect: F&S Partners Inc.; Photo: ©Craig D. Blackman, FAIA.

236

|

ARCHITECTURAL PRECAST CONCRETE

ture panels, quoins, and parapet wall cap copings (Fig.
3.5.106). However, designers should be aware that
most architectural precast concrete producers do not
manufacture cast stone and availability and cost should
be checked prior to specifying.
The high-tech research laboratory reflects the gothic
architecture of its historic campus (Fig. 3.5.107[a]),
characterized by large stone piers and pointed arches.
Buff-colored cast stone with a light sandblast finish
create some of the tower and interest elements of the
building. The cast stone and precast concrete complement other colors and patterns on campus in a more
contemporary and interpretive way (Fig. 3.5.107[b]).

SURFACE AESTHETICS

3.5.14 Architectural Trim Units

3

(a)

Fig. 3.5.107(a) & (b)
Duke University – Center for Models of
Human Disease,
Durham, North Carolina;
Architect: Lord Aeck Sargent;
Photos: Jonathan Hillyer Photography, Inc.

(b)

ARCHITECTURAL PRECAST CONCRETE

|

237

3

SURFACE AESTHETICS

3.5.15 Matching of Precast and Cast-In-Place Concrete

3.5.15 M
 atching of Precast and CastIn-Place Concrete
Architectural precast concrete panels are often used in
combination with architectural cast-in-place concrete.
The exact “matching” of finishes is extremely difficult
and may not be achievable in a cost-effective manner.
The process must be planned prior to start of construction in order to consider adjustments in mixture design,
placement technique, methods of consolidation, and
finishing procedures. Samples and full-scale mockups
should be prepared for both the architectural precast
concrete and the architectural cast-in-place concrete,
and normal differences addressed prior to finalizing
either finish.
Generally, only large projects can justify the increased
cost required for mockups, mixture design control,
tight leakproof forms, and ready-mix concrete using

special or colored cements, or uncommon aggregates.
All of these features may be needed to obtain a satisfactory match-up between architectural cast-in-place
concrete and precast concrete panel finishes.
Aggregate orientation cannot be controlled during
placement and consolidation of the architectural castin-place concrete; rounded or cubical coarse aggregate
are best when trying to match finishes using the different production methods. An acceptable blending
of the two different production techniques should acknowledge the distinction between the processes. It
may be desirable to provide for contrasting colors and
textures between precast concrete and cast-in-place architectural elements. Finishes such as chemical retardation or sandblasting should be used to obtain the best
match. The precast concrete façade in Fig. 3.5.108 has
expressed recesses at exposed cast-in-place concrete
columns and open precast concrete frame work at wall
Fig. 3.5.108
Lumbermens Mutual Casualty Company
Long Grove, Illinois;
Architect: Holabird & Root;
Photo: ©Don DuBroff/Charlotte.

238

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.5.15 Matching of Precast and Cast-In-Place Concrete / 3.5.16 Finishing of Interior Panel Faces

end sections. In tying the precast concrete components
into the columns, the same high level of sandblast finish was required on all exposed surfaces. Bug/blow
holes in cast-in-place concrete are not unusual when
the texture selected is a sandblast finish.
When both techniques meet in the same plane, the
architectural cast-in-place concrete tolerances must be
strictly enforced. Differences in the curing methods
between the two techniques, even with identical mixtures, may cause color variations in the finish, particularly if the precast concrete uses accelerated high-temperature curing. Even when the match-up is very good
at the time of initial construction, different weathering patterns may result from dissimilarity in concrete
densities.

3.5.16 F inishing of Interior
Panel Faces
The back of a precast concrete panel (normally the
face‑up side in the mold) is typically designed to be
flat and level during the casting operation. This is particularly important if this face is exposed as part of the
interior finish.

same appearance.
Vibration (consolidation) of precast concrete usually
forces large aggregate pieces to the bottom of the mold
whereas finer particles are displaced to the top face. To
expose aggregate on the top surface, this surface may
be either water-washed and brushed after the concrete
has obtained initial set, a retarder may be applied and
the surface washed the following day; or the surface
sandblasted in the usual manner. In either case, handplacing (seeding and tamping or rolling) of the larger aggregates should be done after concrete consolidation.
Interior finish requirements should be related to the
configuration of the precast concrete units. This will
be apparent by considering the different mold setups
shown in Fig. 3.5.109. “A” is the most common condition. “B” is similar, but with bulkhead-type top-forms
used to form a beam (or haunch) on the back of the
panel. “C” shows a mold with special top-forms for
molding the interior of the returns.

Fig. 3.5.109 Typical molds and finishes of back of panels.
1

1

A

Exposed interior surfaces should have finishes that are
realistic in terms of exposure, production techniques,
configuration of the precast concrete units, and quality requirements. The finish requirements and area of
exposure for all exposed unformed surfaces should be
shown on the contract drawings. The back of a precast
concrete panel may be given a screed, light broom,
float, trowel, stippled, water-washed or retarded exposed-aggregate finish or sandblast finish. A trowelled
finish is the most common interior finish. Troweling
frequently darkens the surface in uneven patterns.

Typical mold, easy finish of back

B

2

2

1

1

1

1
Concrete will be uneven
under top forms

All edges of precast concrete units that are exposed
to view in the completed structure and caulked should
have a radius, rather than be left as a sharp corner. For
backs of panels that are to receive insulation, the need
for a particular degree of smoothness will depend on
the insulation. The finish should normally be as smooth
as a good wood float finish. If one face of a unit must
be absolutely smooth and true, it should usually be the
face-down surface for uniformity and economy.
The exposed-aggregate procedures for finishes cast
face-up are not similar to those previously described
for face-down casting and may not always result in the

3

Mold with haunch or beam projections on back

Interior finishes will not be
suitable for exposure
3

1

1

3

C

2

2

ARCHITECTURAL PRECAST CONCRETE

|

239

3

SURFACE AESTHETICS

3.5.16 Finishing of Interior Panel Faces / 3.5.17 Acceptability of Appearance

Where top forms are necessary, covering of the back
of panels is recommended, because finishing cost will
be higher than for plain, level surfaces.

panels is normally less with white concrete. In climates
with intermittent dry and wet conditions, drying-out
periods may produce temporary mottled appearances.

The floating or finishing operation should not result
in high areas or ridges around plates or inserts that
have been cast into the unit. Screeded areas should
ensure a uniform thickness across the entire unit that
is within tolerance limits.

It would be optimistic to imagine that every unit cast
during the course of a contract will be identical. The
following is a list of finish defects and/or problems
that are normally unacceptable in high-quality architectural precast concrete. Design and manufacturing procedures should be developed to counteract
or remedy them, unless they are specially desired by
the architect or are inherent in the design of the unit.
If such “defect expressions” are specified by the architect, or are unavoidable, they should be agreed upon
by the precaster and the architect in the form of approved samples and/or initial production units.

3.5.17 A
 cceptability of Appearance
Contract documents must identify who the accepting
authority will be: architect, owner, general contractor,
or site inspector. One person must have final and undisputed authority in matters of acceptability of color,
finish, and texture, in compliance with the contract
documents.
Determining acceptable uniformity of color, finish,
and texture is by visual examination, and is generally a
matter of subjective, individual judgment and interpretation. The acceptable variations should be determined
at the time the visual mockups or initial production
units are approved (see Section 3.2.4). Accordingly, it is
beyond the scope of this manual to establish definitive
rules for product acceptability on the basis of appearance. However, suitable criteria for acceptability require
that the finished concrete face surface should have no
readily visible imperfections other than minimal color
and texture variations from the approved samples or
evidence of repairs when viewed in good, typical daylight illumination with the unaided naked eye at a 20
ft (6 m) or greater viewing distance. Appearance of
the surface should not be evaluated when light is illuminating the surface from an extreme angle as this
tends to accentuate minor surface irregularities due to
shadowing.
Building façades may be oriented such that sunlight
just grazes the surface at a particular time of day. This
causes otherwise imperceptible ripples, projections,
and misalignments on the surface to cast long shadows and be grossly exaggerated in appearance. The
shadows may last briefly. The actual time at which they
appear varies with the season for a particular wall.
Precast concrete, like any building surface, is subject
to manufacturing and alignment tolerances so that the
effect cannot be avoided.
Units should be assessed for appearance during dry
weather. The difference in tone between wet and dry

240

|

ARCHITECTURAL PRECAST CONCRETE

Erected panels not complying with the contract documents may require additional work. The architectural
precast concrete panels should be corrected to match
the repairs demonstrated on the mockup.
1. Ragged or irregular edges. When sharp edges
are specified, minor chips and irregular edges are
unavoidable and should be acceptable.
2. Excessive air voids (commonly called bugholes) evident on the exposed surfaces. If the
air voids are of a reasonable size, 1/8 to 1/4 in. (3 to 6
mm), it is recommended that they be accepted as
part of the texture. Filling and sack-rubbing could
be used to eliminate the voids. However, this procedure may cause color differences. Samples or the
mockup panel should be used to establish acceptable air void size, frequency, and distribution.
3. Adjacent flat and return surfaces with greater
texture and/or color differences than the approved samples or mockups.
4. Casting and/or aggregate segregation lines
evident from different concrete placement
lifts and consolidation.
5. Visible mold joints, seams, or irregular surfaces in excess of or larger than those acceptable
in the approved samples or mockups.
6. Rust stains on exposed surfaces.
7. Excessive variation of texture and/or color
from the approved samples, within the unit,
or compared with adjacent units.
8. Blocking stains evident on exposed surface.
9. Areas where the backup concrete penetrated
through the facing concrete.

SURFACE AESTHETICS

3.5.17 Acceptability of Appearance / 3.5.18 Repair and Patching

10. F oreign material embedded in the face of
the unit.
11. Repairs visible at 20 ft (6 m) or greater viewing
distance.
12. Reinforcement shadow lines (see Section 4.4.5).
13. C
 racks visible at a 20 ft (6 m) or greater viewing distance. The cement-rich film on smooth
concrete may develop a network of fine random
hairline cracks (crazing) when exposed to wetting
and drying cycles. A hairline crack is defined as
a surface crack of minute width and rarely more
than 1/8 in. (3 mm) deep, visible to the naked eye
but not measurable by ordinary means. One of
the primary causes of these types of cracks is
the shrinkage of the surface with respect to the
mass of the unit, due to a cement-rich mixture.
Another cause may be stripping of the panel too
early (inadequate strength and curing), although
this type of crack may be structural.
Crazing is merely a surface phenomenon (penetrates only as deep as the thin layer of cement
paste at the surface of the panel) and has no structural or durability significance but it may become
visually accentuated when the surface is wetted
or dirt settles in these minute cracks. Crazing is
more likely to show up in white or light-colored
concrete than with gray or a dark color, because
the dirt is more visible on the white, but the effect
will depend on the character of the cement film.
A relatively lean, properly consolidated concrete
mixture will show little crazing, in contrast to a
mixture rich in cement and water. Where crazing
occurs, a horizontal surface will be affected more
than a vertical surface due to the settlement of dirt
on the former. Crazing generally will not appear
where the outer cement skin has been removed
by a surface finishing technique. Such cracks are
of little importance and should not constitute a
cause for rejection.
Precast concrete generally undergoes far less
cracking than cast-in-place concrete. This resistance to cracking is due, in part, to the greater
compressive and tensile concrete strengths possible with precast concrete. It should be recognized that a certain amount of crazing or cracking
may occur without having any detrimental effect
on the structural capacity of the member and it is
impractical to impose specifications that prohibit

3

cracking. However, in addition to being unsightly,
cracks are potential locations of concrete deterioration, and should be avoided if possible. Proper
reinforcement locations, prestressing, and proper
handling procedures are the best methods to keep
cracks to a minimum. The acceptability of crazing
or cracking should be determined with respect to
actual service condition requirements, structural
significance, and aesthetics.
Tension cracks are sometimes caused by temporary loads during production, transportation, or
erection of the products. The amount and location of reinforcing steel has a negligible affect on
performance until a crack develops. As flexural
tension increases above the modulus of rupture,
hairline cracks will develop and extend a distance
into the element. If the crack width is narrow, not
over 0.012 in. (0.30 mm), the structural adequacy
of the casting will remain unimpaired, as long as
corrosion of the reinforcement is prevented (see
Section 4.4.7). Accordingly, wall panels containing cracks up to 0.005 in. (0.13 mm) wide for surfaces exposed to the weather and 0.012 in. (0.30
mm) wide for surfaces not exposed to the weather
should be aesthetically acceptable, provided the
reinforcement is galvanized or otherwise corrosion
resistant. The limitation on crack-size specified
is for structural reasons. The aesthetic limitation
will depend on the texture of the surface and the
appearance required. On coarse textured surfaces, such as exposed-aggregate concrete, and on
smooth surfaces comparable to the best cast-inplace structural concrete, the structural limitation
would be aesthetically acceptable. For smooth surfaces of high quality it may be desirable to limit
cracking in interior panels to 0.005 in. (0.13 mm).
In addition, it should be noted that cracks may become even more pronounced on surfaces receiving a sandblasted or acid-etch finish.

3.5.18 Repair and Patching
The techniques and materials for repairing precast
concrete are affected by a variety of factors including
mixture ingredients, final finish, size and location of
damaged area, temperature and humidity conditions,
age of member, and surface texture.
A certain amount of product repair is to be expected
as a routine procedure. The precast concrete elements

ARCHITECTURAL PRECAST CONCRETE

|

241

3

SURFACE AESTHETICS

3.5.18 Repair and Patching / 3.6.1 Weathering General

may sustain superficial damage (minor chipping and
spalling) during handling, transportation, or erection
that may require jobsite repair. Repair and patching
of precast concrete is both an art and a skill requiring
expert craftsmanship and careful selection and mixing of materials, to ensure the end result is structurally sound, durable, and aesthetically pleasing. Major
repairs should not be attempted until an engineering
evaluation is made to determine whether the unit will
be structurally sound and if so determined, the repair
procedure should be approved by the precast concrete
design engineer.
Trial mixtures are essential to determine exact quantities for the repair mixture to effectively match color
and texture. This is best determined by applying trial
repairs to the project mockup or small sample panels
[12 x 12 in. (300 x 300 mm)]. The trial repairs should
be allowed to cure, followed by a normal drying period. This is important because curing, weathering,
and ultraviolet bleaching of the cement skin affect the
finished color. Because of these factors, it is unreasonable to require an instant color match at time of repair.
All parties should agree on the acceptability criteria for
repairs at an early stage of the project.
It is recommended that the precaster execute all repairs
or approve the methods proposed for such repairs by
other qualified personnel. The decision of when to perform the repairs should be left up to the precaster, who
should be responsible for satisfactory final appearance.
It is important that all repair of damaged precast concrete unit edges be carried out well in advance of the
joint sealing operation. The repair work should be fully
cured, clean, and dry prior to caulking.
Repairs should be done only when conditions exist
that ensure that the repaired area will conform to the
balance of the work with respect to appearance, structural adequacy, and durability. Slight color variations
can be expected between the repaired area and the
original surface due to the different age and curing
conditions of the repair. With time (several weeks) and
exposure to the environment the repair will blend into
the rest of the member so that it becomes less noticeable. Excessive variation in color and texture of repairs
from the surrounding surfaces may be cause for rejection until the variation is minimized. Repairs should be
evaluated when the concrete surface is dry.
Precise methods for repair are not detailed in this manual. (See PCI Erectors Manual: Standards and Guidelines
for the Erection of Precast Concrete Products, Appendix

242

|

ARCHITECTURAL PRECAST CONCRETE

E-Repair Procedures, MNL-127 for guidance on repair
techniques and materials.) Precaster should be requested to submit recommended repair procedures.

3.6 WEATHERING
3.6.1 General
A primary consideration in the architectural design of buildings should be weathering, that is, how
the appearance changes with the passage of time.
Weathering affects all exposed surfaces and cannot
be ignored. The action of weathering may enhance
or detract from the visual appearance of a building,
or may have only a slight effect. The final measure of
weathering’s effect is the degree to which it changes
the original building appearance and distorts the designer’s original intention by streaking or shading.
Visual changes occur when dirt or air pollutants combine with wind and rain to interact with the wall materials. The run-off water may become unevenly concentrated because of façade geometry and details. The
manner in which water is shed off the structure depends primarily on the sectional profiles of the vertical
and horizontal discontinuities designed into the wall.
Through the years, designers controlled the water
flow down specific parts of a structure with copings,
drip grooves, gargoyles, window sills, and plinth details. However, many of these useful and relevant details have been discarded as superfluous decoration.
For architectural precast concrete (as well as all other building materials), the awareness of weathering
should be reflected in the design of wall elements and
the integration of windows to control water sheeting
and penetration and to manage water run-off. Staining
that occurs through differential surface absorption and
uneven concentrations of dirt due to water run-off are
considered the most common weathering problems.
Many of the effects of weathering can be predicted
by studying local conditions and/or existing buildings
in the area. This will often give a clear indication of the
levels of pollution; the velocity, principal direction and
frequency of wind; and the intensity, duration, and
frequency of rainfall; together with records of temperatures and relative humidity. All these factors will
affect the way exposed concrete will get wet and dry
out. With proper attention to the causes and effects
of weathering, potentially detrimental results can be
eliminated or at least minimized. Design tools for con-

SURFACE AESTHETICS

3.6.1 Weathering General

trol of weathering are the massing and detailing of the
building and the color, texture, and quality of the surface finishes. Precast concrete will become dirty when
exposed to the atmosphere, just like any other material. Fortunately, with architectural precast concrete,
the designer can choose shapes, textures, and details
to counteract many of the negative effects of weathering. Although regular cleaning of a building may make
detailing a less critical factor, maintenance costs should
be balanced against initial design costs.
One of the major contributing factors to the weathering of precast concrete is dirt in the atmosphere.
Atmospheric dirt or air pollutants include smoke or
other gases, liquid droplets, grit, ash, soot, organic
tars, and dust. Gaseous pollutants include SO2, NOX,
H2S, NH3, and O3. Sulphur dioxide (SO2) can react with
the lime in the concrete and the oxygen from the air
to form gypsum (see Section 5.2.4). Gypsum’s solubility allows for it to be washed away, taking dirt with it.
Where there is insufficient water to wash it away it can
encapsulate dirt and hold it.
The concentration of SO2 and other corrosive compounds is high in some urban environments. When
dissolved in rainwater, SO2 produces dilute sulfurous
or sulfuric acid. These acids etch cement-rich paste
and carbonated precast concrete surfaces, producing a
gradual change in color as the fine aggregate becomes
exposed.

3

Figure 3.6.1 shows the pH of acid deposition falling in the U.S. during 2005. Although acid deposition
(acid rain) is technically defined as precipitation with
a pH level below 5.6, some researchers believe that it
should be defined as low as 5.0. Using either definition, acid deposition has a far-reaching impact on both
the U.S. and Canada.
In areas with unusually high concentrations of corrosive elements (pH of rainwater lower than 5.0), the
designer should detail the façade for water run-off,
specify concrete strengths and durabilities normally
associated with architectural precast concrete, provide
the required cover over reinforcement, avoid soft aggregates such as limestone and marble, and suggest
more frequent washings of the building.
The flow of rainwater across the building’s façade
has a profound affect on weathering patterns because
rain run-off redistributes particulate matter that has
been deposited fairly uniformly on the external wall
surfaces. This deposit takes place more rapidly on surfaces facing upward and also on surfaces with a coarse
texture. The designer should attempt to anticipate and
plan for water flow over the wall, tracing water flow
to the final drainage point or to ground level, particularly where discontinuities exist. When run-off reaches
a discontinuity the water may bead and drip free. This
may increase or decrease the run-off concentration,
affecting both the run-off’s ability to carry suspended

Fig. 3.6.1 Hydrogen ion concentration as pH from measurements made at the Central Analytical Laboratory, 2005 — distribution of
acid rain.

ab pH

Sites not pictured
AK03
5.3
VI01
5.1

Source: National Atmospheric Deposition Program (NRSP-3),
2007. NADP Program Office, Illinois State Water Survey,
2204 Griffith Dr., Champaign, IL 61820.

5.3
5.2 - 5.3
5.1 - 5.2
5.0 - 5.1
4.9 - 5.0
4.8 - 4.9
4.7 - 4.8
4.6 - 4.7
4.5 - 4.6
4.4 - 4.5
4.3 - 4.4
< 4.3

ARCHITECTURAL PRECAST CONCRETE

|

243

3

SURFACE AESTHETICS
3.6.1 Weathering General

dirt particles, and the subsequent drying behavior of
the wall. Such changes of flow concentration may disfigure the building surfaces.
Rain is primarily a cleansing agent for building surfaces. However, at some stage the water will also pick
up particulate matter already deposited on the walls
and it becomes a soiling agent. The preferred lines of
water flow must be arranged through shaping of surfaces and textures so that, at the point where water is
expected to become a soiling agent, it will not detract
from the finishes or forms of the building elements.
Particulate matter will drop out of the run-off water
when water flow velocity is decreased; for example,
when the run-off is allowed to fan out. It may be necessary to have frequent details to throw water clear
of the building, collect the water, or spread the water
uniformly across sloped surfaces. Such details should
be continuous to prevent differential rainwashing, or
must terminate at bold vertical features, or maintain
a clear distinction between washed and soiled areas.
These differences can then, if required, be emphasized
by the use of varying surface finishes.
The migration of run-off water is affected by:
1. The location and concentration of rain deposits.
2. The properties of water in contact with materials,
especially surface tension.
3. The forces of wind and gravity.
4. The geometry, absorption, and texture of the building surface.
5. Drips.
The amount of rainwater, and the velocity and angle
at which it falls is markedly different on each side of
a building and at different heights. Therefore, it is not
reasonable to expect equal weathering of all parts of
the building. The influence of tall or massive buildings,
projections, courts, or passages on prevailing winds
can cause wind eddies to upset the natural flow of air
and rain. This makes the effect of rainwater very difficult to predict. Also, a wall that receives a great deal
of sunlight will dry out a lot faster and will be less likely
to attract airborne particles.
The wettest portion of a building is typically the top
corners of the windward face, followed by the top and
side edges. The side wall, which is parallel to the wind
direction, remains relatively dry. A wide face remains
drier overall, particularly in its center region, relative to
a narrow face. The taller the building and the higher

244

|

ARCHITECTURAL PRECAST CONCRETE

the wind velocity, the relatively narrower the band of
high rain impact. Increased wind speed also appears
to cause greater wetting in the center of a façade.
Corners may be subjected to 20 to 30 times more rain
impact as compared with the central region of the
building face.
A large roof overhang allows the wall surface to
weather uniformly. Peaked roofs, cornices, or horizontal projections can substantially reduce the amount
of rain that falls on a façade by reducing lateral acceleration at the wall-roof intersection. The horizontal
projection should have a minimum projection of 12 in.
(300 mm) to throw water off. In many cases, in order
to be in scale with the building it will be much more.
The projection must have a drip on the underside to
prevent water running back across the soffit. This stops
soffit staining, and also prevents random staining on
the surface below.
During storms, driving rain can come from any direction, but the quantity of water available on a façade
for washing is normally determined by its relationship
to the prevailing wind and the intensity of rain from
that direction. Wind movements around buildings are
affected not only by major climatic factors but by local
topography, adjacent buildings, and groups of trees.
All these will affect the amount and position of driving
rain hitting a building and the way water runs down
the façade. The drops of driving rain are guided for
the most part by the air currents around the building
and external wall components. The pattern of these air
currents is independent of building height. Small obstacles give rise to sudden changes in direction of the
air current and the raindrops cannot follow these sudden changes. The mass forces carry them forward to
the obstacle. On one- or two-story buildings, the driving rain reaches the lower parts of the walls. Dirt stain
patterns do not usually occur on such low buildings.
Air currents against buildings taller than a couple of
stories are, on the other hand, deflected so gently that
the air has time to re-orient the raindrops. When the
wind blows at a building, some of the air will rise to
pass at an increased velocity over the top; the rest will
form a horizontal vortex and spiral away around the
ends (Fig. 3.6.2). Less than half the quantity of rain
that should pass through a free air cross-section of
the same size as the building is caught by an external
wall. This applies regardless of the wind force. The rain
mainly strikes the top parts of the wall. Only edge sections (corners) are reached by the driving rain and in

SURFACE AESTHETICS

3.6.1 Weathering General

Wind
direction
15 stories

3 stories

Fig. 3.6.2
Wind movement around building and rain wetting of façades.

the central sections the raindrops move almost completely parallel to the wall. As a result, water run-off
very seldom reaches all the way down to the ground,
except at corner areas and projecting components, unless the duration of the rain is quite long. Therefore,
special care should be taken to ensure that water is not
allowed to run down surfaces unless there is enough
water to wash the surfaces completely. When the runFig. 3.6.3

3

off water reaches the area of wall that is protected
from driven rain by the horizontal vortex it will be absorbed into the surface causing a typical zigzag dirt
line. The level at which the jagged line of dirt forms
will be governed by a combination of the height of the
vortex and the absorbency of the precast concrete. The
height of the vortex above the ground is determined by
the height of adjacent buildings or other obstructions
over which the wind has passed. A typical weathering
pattern caused by rain and prevailing wind is illustrated
in Fig. 3.6.3. Parts of the building façade are clean in
areas where it is washed by rain, even though the remainder of the building has become soiled.
Rounded or splayed corners reduce wind speed at
the edges of buildings and may be useful to avoid the
heavy concentrations of driving rain that are typical of
these locations. Also, continuity of water flow between
surfaces is improved when corners between them have
rounded instead of sharp edges. A joint, groove, or
projection near a corner with a long return should be
used to catch the rainwater and prevent partial dirt
washing from water blown around the corner.
The raindrops that reach a wall surface are absorbed
to different extents depending on absorption and
moisture content of the wall material. Precast concrete normally has a medium to low water absorbency.
Water run-off on concrete surfaces consists of a very
thin layer, 0.01 in. (0.25 mm) thick, and only occurs
if the absorption of the concrete is lower than a certain value. The run-off flows slowly (up to 3 ft/min.
[0.9 m/min.]) and vertically down the wall with lateral
winds having an insignificant influence. When the water reaches lower sections, which have been struck by
less driving rain and are drier, it is absorbed. The dirt
accompanying the water is deposited in new places,
unevenly soiling the surface. Also, a façade with high
absorption normally becomes wet rapidly and remains
damp for a longer period than a façade with low absorption. Airborne dirt (soiling particles) adheres easily
to high absorption concrete. It is desirable to break up
large areas of concrete, extending over several stories,
with horizontal features that either collect or throw off
the water at intermediate positions. These features will
reduce the amount of water on the surface, reduce the
differences between panels at different levels on the
façade, make the change from washed to unwashed
into a gradation instead of a clearly visible line and by
producing interest and shadows will make any changes less noticeable.

ARCHITECTURAL PRECAST CONCRETE

|

245

3

SURFACE AESTHETICS
3.6.1 Weathering General

each window—the very thing the designer must guard
against if differential patterning is to be avoided. This
flow must be dissipated, breaking up its concentration.
Furthermore, there is always a tendency for water flow
to be in greater volume at the edges of the glass (the
smallest amount of wind tends to drive rain toward
the edges of the glass). Figures 3.6.4(a) and (b) show
a soiling pattern caused by water run-off carrying particulate matter down the mullion and over the precast
concrete. In Fig. 3.6.4(a), an attempt to minimize staining resulted in grooves being cut under the mullions.

(a)

Fig. 3.6.4(a) & (b)Water flow over glass depositing dirt.

(b)

The water run-off on concrete surfaces has a tendency to divide into separate streams determined by
microscopic irregularities or differences in absorption
of the surface when the water layer thickness decreases below a critical value. This breakdown into irregular, separate streams takes place mainly on smooth
or lightly textured surfaces but can also occur on surfaces with exposed aggregates. However, a uniformly
distributed, broken flow is more likely to occur over
heavily textured materials (Fig. 3.6.5). These streams
recur at the same locations during most rainfalls and
are reflected in the soiling pattern. The streams of water broaden out laterally when they meet horizontal or
moderately sloping obstacles. They also follow surfaces facing downward (horizontal surfaces) in a similar
manner. Consequently, the design of drips is extremely
important. Surface tension allows flows to take place
along the underside of horizontal surfaces. Therefore,
horizontal ledges, returns, and bullnoses should have
a drip groove in the underside to prevent water running back onto the façade and causing staining. The
path followed by the water from the lowest points of
Fig. 3.6.5 Water flow over smooth versus rough texture.

Surface tension causes droplets of water to coalesce
on non-porous surfaces such as glass and metal and to
drain in irregular streams. Glass areas cause build-up
of water flow. Because glass is a non-absorbent material, the flow rate of water down its surface is fast,
and there is little time lag in its throw-off. By contrast,
rainwater flowing down an adjacent concrete wall
surface will be slower (depending on the surface texture) and its throw-off will be less complete. As a result, there is a concentration of water at the bottom of

246

|

ARCHITECTURAL PRECAST CONCRETE

SURFACE AESTHETICS

3.6.1 Weathering General

Fig. 3.6.6 Volume of rain likely to hit wall surfaces.

In A and B lengths “a” and “b”
are directly related to the volume
of first contact rain

a
A

Rain

Rain
b

Vertical surface

Dirt becomes
resistant to
clean run-off

b

Best slope for
self-cleaning

B
Steep forward-sloping
surface

Flat forward-sloping surface

Rain
Backward-sloping surface
In C the surface is completely
shaded from first contact rain

C

Rain is assumed as 10° to the vertical
for diagrammatic purposes

these drips should also be taken into consideration in
the design.
A coarse texture often stains in a way that is not too
objectionable. The naturally present surface contrasts
are usually enhanced.
The vertical angle of a surface has a major influence
on the quantity of pollutants it collects and how they
are discharged during rain. Figure 3.6.6 shows the volume of rain assumed to hit a building surface depending on the orientation of the surface. For diagrammatic
purposes, the angle of rainfall direction is assumed to
be 10° from the vertical. However, the variability of rain
under actual conditions makes all but a general prediction difficult. Vertical or near-vertical surfaces receive
insufficient rainwater to be self-cleansing. Steep forward-sloping surfaces usually weather cleaner. Large
areas may begin to collect dirt at the lower end unless
the angle is steep. With heavy rain, the dirt on horizontal surfaces and surfaces that have little slope may
be partially washed off, streaking the surfaces below.
In the case of light rain or drizzle, the dirt may collect
and slowly flow down other surfaces in the general

3

direction of the water flow, resulting in pronounced,
random streaking. Backward-sloping surfaces collect
little or no rain but are likely to be subject to a partial, nonuniform water flow from above, which may
carry dirt and cause serious streaking. Backward-sloping surfaces are often seen in shadow. In this case, the
accumulation of dirt is not particularly noticeable if the
dirt is acquired evenly without disfiguring streaks. The
following guidelines are derived from the weathering
of exposed surfaces:
• Avoid horizontal planes—they collect the most dirt
and are the most exposed to rain.
• Use sloping rather than flat upper surfaces—the
advantage of a sloping upper surface is that the
dirty water is immediately drained away. It is best to
adopt a steep slope and a limited height to encourage fast run-off of rain and complete washing. If
there is a risk of an uneven staining pattern, then
opt for a darker color and/or a complex surface texture, for example, deep grooves.
• Avoid run-off into the elements in the rain shadow.
• Avoid run-off from horizontal or gently sloping upper surfaces on vertical sections of walls. Investigate
whether the combination of the two surfaces cannot be replaced by a slope. The pattern of the spandrel under the windows is influenced by this.
• Every inclined or vertical surface with underside
edges must be provided with a well-profiled drip in
order to prevent the rain running off into the rainshadow area.
The façade geometry of buildings is usually responsible for local concentrations of run-off. Such concentrations lead to the characteristic marking patterns
frequently observed on building surfaces: dirt accumulation, dirt washing, and white washing (Fig. 3.6.7).
Fig. 3.6.7 Characteristic marking patterns.

DW
DA
WW

DW — Dirt-washing
WW — White-washing
DA — Dirt-accumulation

ARCHITECTURAL PRECAST CONCRETE

|

247

3

SURFACE AESTHETICS

3.6.1 Weathering General / 3.6.2 Surface Finish

New buildings may show dirt washing at locations
of concentrated run-off while their over-all surfaces
are still quite clean. Later, the same areas may exhibit
white washing (lighter, cleaner streaks) after adjacent
surfaces have been darkened by dirt accumulation.
Adjacent precast concrete units should have faces
aligned within accepted industry tolerances (see Section
4.6.3). Any discrepancy may pass undetected on a new
building, but weathering will eventually emphasize the
offset with uneven staining of adjacent units.
When mullions and other vertical elements meet a
horizontal element, the configuration often causes a
concentration of flow and uneven weathering. Vertical
surfaces can be protected and weathering minimized
by providing steeply sloping overhangs with drips.
These tend to reduce dirt accumulation and the washing of dirt onto the vertical surfaces below.
The intersection of horizontal and vertical projecting
elements almost always creates dirt streaks. Such streaks
run back from the edge of exposed columns and below
the ends of horizontal elements even when they are
steeply sloped at the top surface. To avoid such streaks,
the horizontal element should be stopped short of the
column. This confines rain run-off to the horizontal element and permits unimpeded washing of the column.
Channeling of the column faces also will help prevent
water from running back along the edges.
Water flowing laterally or diagonally downward on
a surface will concentrate where it encounters vertical projections or recesses. The secondary airflow due
to wind is also important. It concentrates run-off at
the outside corners of the building, at columns, and at
inside corners of vertical projections. Surface tension
contributes to this effect by preventing flow back from
vertical edges of small elements such as window mullions, often concentrating the flow at the corners.
In areas where nearby buildings show the undesirable
effects of weathering and the local atmosphere is laden
with pollutants, it is recommended that consideration
be given to the use of sealers to increase rain run-off,
reduce surface absorption of the concrete, and facilitate cleaning of the surface (see Section 3.6.6).
Careful attention should be given to the effects of
exterior details on the weathering pattern of the building. Appropriate design details help to avoid many of
the more unsightly effects of dirt streaking and differential washing.

248

|

ARCHITECTURAL PRECAST CONCRETE

Thus, the principles the designer should consider to
minimize the visual effects of water flow are:
• Provide steeply sloping surfaces to allow self cleaning, and limit the distance of water flow.
• For vertical surfaces, detail surface finishes that disperse rainwater flow over the surface (for example,
exposed aggregate) or provide vertical striations
that direct the flow more evenly.
• Avoid concentration of water flow.
• Prevent water flow over sheltered positions by detailing drip grooves to throw the water clear.

3.6.2 Surface Finish
Concrete surface finishes vary considerably in their
ability to take up and release dirt under weathering
conditions. They should therefore be chosen for their
so-called “self-cleansing” properties. But the selection
of color and texture may have an aesthetic significance
greater than the effects of weathering. The economics
of varying the surface finish from one part of the building to another should be investigated as the weathering characteristics will be different.
The surface of smooth precast concrete is hard and
impervious and easily streaked by rain, unless there is
enough water to form a complete film on the surface.
Weathering patterns are determined by the shape and
smoothness of the units and joints, which are particularly vulnerable. Any irregularity in a smooth surface
will be exaggerated by weathering patterns. Non-repeating, irregular, and concentrated streams tend to
form on smooth or lightly textured materials. Lighttoned and smooth surfaces accentuate the contrast
between washed and unwashed areas.
Textured finishes accumulate more dirt, but they can
maintain a satisfactory appearance. The aggregate
tends to break up and distribute water run-off more
evenly, reducing the streaking that appears on smooth
surfaces (Fig. 3.6.8). Textured finishes also have a slower drain-off because each stream is small. The irregularities and shadows on the surface also tend to mask
discoloration. It is not reasonable, however, to expect
an exposed-aggregate finish to deal with all problems
of weathering. The way water moves on such a surface
is different, but concentrated flows or their effects will
still be visible and must be controlled.
Rounded aggregates are preferred because they tend
to collect less dirt than angular aggregates with rough

SURFACE AESTHETICS

3.6.2 Surface Finish

3

Fig. 3.6.8 Streaking on smooth versus textured finishes. Photo: Bill Rothschild.

texture. However, dirt pickup is generally confined to
the matrix. For this reason, as well as for architectural
appearance, the area of exposed matrix between aggregate particles should be minimized. The smooth,
nonporous surfaces of the aggregates allow less dirt to
deposit and promote more run-off to increase washing
of the surface. At the same time, a slightly recessed
or a darker matrix helps to absorb and mask pollution
deposition.
Extreme color differences between aggregates and
matrix will create uniformity problems. For example,
large exposed aggregates of light color provide heavily
textured surfaces that may seem to be very dirty with
time because the matrix becomes very dark and the
high spots of the aggregates are washed clean. In some

cases, uniformly colored light surfaces contrasted with
uniformly dark-colored surfaces may be used to accentuate the depth of relief on a building face. Smooth
units made with dark-colored sands will slowly become
darker with age when subject to weathering because
the surface film of cement paste erodes away, exposing the sand. Therefore, wide differences between the
color of the cement and the sand should be avoided.
The use of appropriate colors and textured surfaces
can help to mask the effect of dirt deposits. The overall
darkening in tone that takes place is unlikely to be objectionable unless streaking occurs. Medium textured
finishes may still allow water to run or be wind-driven
into streams, causing irregular streaks. Vertical ribs or
flutes that help the designer give expression to a fa-

ARCHITECTURAL PRECAST CONCRETE

|

249

3

SURFACE AESTHETICS

3.6.2 Surface Finish / 3.6.4 Efflorescence on Precast Concrete

Fig. 3.6.9
Arizona Public Service Administration Complex, Phoenix, Arizona;
Architect: DFD Cornoyer-Hedrick formerly Comoyer-Hedrick Architects & Planners;
Photo: DFD Comoyer Hedrick.

çade will also help to control the run-off and prevent
it from spreading horizontally. As dirt collects in the
hollows, it emphasizes the shadow and, therefore, the
texture itself. The rib must not be too wide otherwise
a soiled pattern may develop in the middle area of the
rib’s upper surface. If the ribs are terminated above the
lower edges of the walls, streaking below the ribs may
occur depending on the depth of projection and the
wind force and direction. As water reaches the bottom edge of a vertical or inclined panel, surface-tension effects cause it to slow down before dripping clear
and it tends to deposit any dirt it has been carrying.
Horizontal ribs or flutes spread stains rather than prevent them and can be used to protect the underlying
surface by deflecting the flow of water. Water flows
on diagonal ribs create a weathering pattern difficult
to predict.

3.6.3 D
 eposits from an Adjacent Surface
or Material
A partial water flow may be stopped from running
over backward slopes by water drips. If dirty water,
thus directed, falls partially onto other surfaces the
problem may be merely relocated. The effect of dirt
washed onto precast concrete surfaces may be diminished by rough textured surfaces with reasonable dark

250

|

ARCHITECTURAL PRECAST CONCRETE

color tones to minimize the visual appearance of the
dirt. Alternatively, the dirty water may be directed off
the building. An example of this is shown in Fig. 3.6.9.
The three-story panels have spandrel areas below windows that project out from the building at each successive level.
Water flowing over copper, bronze, weathering steels,
some silicone sealants, sheet metal flashing, or aluminum, which subsequently flows over concrete, mainly
creates green, rust-brown, or black stains over a period
of time, (see Section 5.2.3). Consideration should be
given to using parapet flashing with a drip detail and
to possibly protecting against corrosion. These types of
discoloring are more difficult to remove than ordinary
climatic dirt. Also, maintenance procedures such as
window cleaning can produce dirt markings on precast
concrete unless care is exercised.

3.6.4 Efflorescence on Precast Concrete
Efflorescence is a frequent issue that the architectural
precaster has to address. Efflorescence often appears
within the first year after the structure is completed—
when the architect, owner, and contractor are most
concerned with the appearance of the new structure.
However, the length of time that lapses before efflorescence occurs may vary greatly; it may appear after

SURFACE AESTHETICS

3.6.4 Efflorescence on Precast Concrete / 3.6.4.1 What is efflorescence

only one day, or within a few days of product stripping, and sometimes not until weeks, months, or (in
rare cases) years have passed. Though generally harmless from a structural viewpoint, the initial appearance
of efflorescence can be detrimental to the appearance
of a finished structure, but should not cause concern.
Recurrent efflorescence, on the other hand, indicates a
chronic moisture problem, and efforts should be taken
to prevent and eliminate it.
The phenomenon of efflorescence apparently does
not follow a firm principle but frequently occurs at
random. Sometimes no problems may be experienced
with one batch of concrete while the next batch
shows a strong tendency to develop efflorescence.
Nevertheless, it must be recognized that, under certain
conditions of construction and exposure, efflorescence
is inherent and unavoidable.
While little attention is given to efflorescence on
white or light-colored surfaces, the contrast on darkcolored concrete is obvious and attention-getting.

3.6.4.1 W
 hat is efflorescence
Efflorescence usually occurs due to the presence of
soluble substances in the materials used to produce
concrete. Water-soluble salts present in concrete materials as only a few tenths of 1% are sufficient to
cause efflorescence when leached and concentrated
at some point on the surface. The amount and character of deposits vary according to the nature of the
soluble materials and the atmospheric conditions. The
chemical components of efflorescent salts are usually
alkali metal and alkaline earth sulfates, hydroxides,
and carbonates. The most common salts found in efflorescence are sodium and potassium sulfate, sodium
bicarbonate, and calcium hydroxide or carbonate. The
sulfates and bicarbonate are readily soluble in water,
while calcium carbonate is not.
At early ages, the hydrated cement in concrete contains a substantial amount (15% by volume of the cement paste) of calcium hydroxide as a normal product
of the hydration reaction between cement and water.
Some of the calcium hydroxide dissolves in the mixing
water, migrates to the surface of the fresh concrete,
and precipitates there. The solubility of the calcium
hydroxide increases with decreasing temperature, and
the greater the solubility, the greater the likelihood of
efflorescence. It is not the water in the concrete that
migrates to the surface with the calcium hydroxide;

3

rather, the calcium hydroxide in aqueous solution
moves through the capillary pores of the concrete to
the surface, where it reacts with carbon dioxide in
the air to form water-insoluble calcium carbonate,
which then appears as a whitish deposit (primary efflorescence). Primary efflorescence can only occur if
the concrete capillaries are filled to the top with water. Precipitation of the calcium carbonate reduces the
local calcium hydroxide concentration at the concrete
surface, thus creating a concentration difference in relation to the interior of the capillary system. New calcium hydroxide is then supplied to the surface from the
interior concrete capillary system. Calcium hydroxide is
much more soluble in water at cold temperatures than
at warm temperatures, such deposits are more common in damp, winter months.
In a subsequent, slower reaction over a concrete age
of one to three years, calcium carbonate can react with
additional carbon dioxide and water to form calcium
hydrogen carbonate (calcium bicarbonate), which is
water-soluble. This type of efflorescence can be partially washed away by rain. The acid constituents of
the atmosphere (for example, sulfur dioxide) can result
in transformation of the calcium hydroxide deposits on
the concrete surface to calcium sulfate. Efflorescence,
therefore, disappears faster in areas with acid rain than
in the those of marine and mountain climates.
Other causes for the migration of calcium hydroxide
to the concrete surface include rainwater, which penetrates or comes in contact with the concrete, or water
of condensation, which may occur within or on the
concrete. Such water is initially free of dissolved calcium ions, but as a result of a concentration gradient,
the calcium hydroxide migrates out of the concrete to
the water on the surface, where it eventually reacts
with carbon dioxide. Such efflorescence, which can occur during the continual curing of hardened concrete,
is referred to as secondary efflorescence. Secondary efflorescence will not usually occur if the concrete surface
has had time to carbonate to any appreciable depth.
No clear distinction can be made between primary
and secondary efflorescence, particularly as the transition between the two is not clear.
Identification of efflorescence deposits is sometimes
useful. X-ray diffraction analysis, petrographic analysis,
and chemical analysis are techniques some commercial
laboratories use to identify efflorescence deposits. In
some instances, it is useful to know both the type of

ARCHITECTURAL PRECAST CONCRETE

|

251

3

SURFACE AESTHETICS

3.6.4.1 What is efflorescence / 3.6.4.2 What causes efflorescence

salt present and its relative quantity. For example, very
soluble salts such as alkali sulfates and bicarbonates
indicate possible material problems, whereas relatively
insoluble efflorescence deposits may indicate problems
related to too much moisture moving into the concrete
or unfavorable curing conditions.
Efflorescence may also occur on any concrete surface
due to migration of salts from the ground up the exterior surface of foundation walls. In the western U.S.,
sodium sulfate in the soil may climb a few inches up
walls and across horizontal surfaces. It may then enter the concrete a short distance and undergo volume
changes that cause loss of concrete surface.

3.6.4.2 W
 hat causes efflorescence
A combination of circumstances causes most efflorescence. First, soluble substances must be in one or
more of the concrete materials. An abundance of calcium hydroxide is always present in concrete, but the
quantity of soluble alkali metals will vary depending
on the cement source. Second, moisture must be present to dissolve the substances. Third, evaporation, hydrostatic, or osmotic pressure must cause the solution
to move toward the surface. And fourth, the solution
must evaporate to leave these substances behind as
efflorescence. If any one of these conditions is eliminated, efflorescence will not occur.
The efflorescence tendency of most high-quality concrete decreases with increasing age and with rapid drying and carbonation of the surfaces. However, concrete
that is constantly or frequently saturated with water
can continue to develop efflorescence for years.
Many factors affecting the occurrence of efflorescence on concrete surfaces interact with one another.
These factors are relative humidity, temperature, and
air movement (these affect the rate of surface drying),
and permeability and texture of the concrete surface.
Some types of efflorescence occur most frequently
when temperatures are low and humidity is high because calcium hydroxide has greater solubility at low
temperatures. In the northern U.S., this happens most
often in the early spring or fall when there are intermittent rains, and the temperatures are still only in the
range of 30 to 50 °F (0 to 10 °C). These conditions
rarely occur in some southern regions; consequently
these regions rarely, if ever, have any problems with
such efflorescence.

252

|

ARCHITECTURAL PRECAST CONCRETE

Several researchers have observed a correlation between weather variables and efflorescence. The research suggests that efflorescence may be related to
climatological data on precipitation, wind, and temperature. Wet, cool, windy weather seems to induce
efflorescence. For 246 U.S. weather stations, an efflorescence index was calculated for each month as
the product of the mean normal precipitation (in.), the
average wind speed (mph), and the reciprocal of the
average, normal, daily mean temperature (ºF). Those
monthly values were then averaged for the year and
are plotted in Fig. 3.6.10 with isopleths.

Fig. 3.6.10 The efflorescence index.

“Masonry’s Latent Defect,” by Clayford T. Grimm, The Construction
Specifier, vol. 51, no. 11, October 1998, page 58.

Efflorescence is more prevalent where the efflorescence index is greater than 0.5. Efflorescence causes
some concern at mid-continent of the United States
and becomes progressively troublesome toward the
east coast. There is less efflorescence from mid-continent to the west coast, except for the far northwest.
High evaporation rates always reduce the degree of
efflorescence, whereas low evaporation rates need not
necessarily result in high efflorescence levels. In the
summer, even after long rainy periods, moisture evaporates so quickly that comparatively small amounts of
salt are brought to the surface. Usually efflorescence is
more common in the fall or early spring when a slower
rate of evaporation allows a greater amount of salts or
calcium hydroxide to migrate to the concrete surface.
Imagine fresh concrete as a material riddled with capillary pores that are filled with an aqueous solution of
the water-soluble components of the cement—mainly

SURFACE AESTHETICS

3.6.4.2 What causes efflorescence / 3.6.4.3 Minimizing efflorescence

calcium hydroxide and alkali sulfates. As the concrete
hardens, the calcium hydroxide located at the surface
of the capillaries reacts with atmospheric carbon dioxide to form calcium carbonate. Due to the formation
of calcium carbonate, the concentration of calcium hydroxide at the mouth of a capillary is lower than inside
it. For this reason, calcium hydroxide is continuously
diffused from the lower layers of the concrete to the
surface. It is true that visible efflorescence can only occur if the capillaries in the concrete are wet throughout.
Only then can the calcium hydroxide and alkali salts
reach the surface. The capillaries are gradually blocked
with calcium carbonate, and the whole process normally comes to a halt. However, if the surface of the
concrete is covered with a film of condensation, the
calcium hydroxide can spread over the entire surface
area and react to form a layer of calcium carbonate,
which is insoluble in water. In this case, efflorescence
will be more severe than when no water film is present
on the surface of the concrete and calcium carbonate
is only found at the capillary surface.
Well designed and produced concretes contain capillary systems in which water not needed for hydration
of the cement paste can make its way (by diffusion)
into the atmosphere. The size of these capillary pores is
of decisive importance with regard to the formation of
efflorescence. The finer and more elaborately branched
the capillary pore system is, the more intensively the
diffusible water is held back by capillary forces and
the drier the ambient air must be to induce the water to evaporate from the pores. Also, the smaller the
capillary diameter, the more rapidly the outlets of the
capillaries at the surface of the concrete are blocked
by calcium carbonate from the reaction of the calcium
hydroxide with carbon dioxide in the air. The formation of efflorescence is stopped as the concrete surface
becomes denser and less permeable.
If the capillary pores in hardened concrete could be
limited to a minimum, this would be a step forward
in solving the efflorescence problem. When concrete
is thoroughly hydrated, its pore structure changes and
its permeability is decreased dramatically, so that, as a
rule, it will not be affected by efflorescence.
Secondary efflorescence can appear on the surface
of concrete during weathering, even if the concrete
has cured properly. What is remarkable is that secondary efflorescence usually occurs for roughly as long
as there is a marked increase in the strength of the
concrete during exposure to moisture. It is therefore

3

likely to be a consequence of further hydration of the
cement. During this process, calcium hydroxide is deposited on the surface and then reacts with carbon
dioxide. Secondary efflorescence appears to reach a
maximum within a year. The maturity of the concrete,
defined as the product of the curing temperature in
degrees Celsius (ºC) and curing time in hours (h), has
been related to secondary efflorescence. At maturity
values above 1300 (ºC × h), efflorescence is usually
minimal.
The more common causes of moderate amounts of
secondary efflorescence do not lie in the transport processes from deeper layers of concrete to the surface,
but are thought to be more closely related to localized conditions on the surface. This is backed up by the
point that the efflorescence would otherwise have to
be very pronounced since there is a much larger quantity of calcium hydroxide available in the concrete as a
whole. When the surface of concrete is sandblasted,
retarded, or acid-etched, the quantity of secondary efflorescence can be greater than it would have been for
a densely formed or finished surface under the same
environmental conditions.
The weathering away of a thin layer of secondary
efflorescence in one to two years can be due to the
slow formation of water-soluble calcium hydrogen carbonate from calcium carbonate. Once this occurs, it
is easily washed away by rain. In areas with little rain
(Arizona, for example), secondary efflorescence is particularly long-lasting. A great deal of rain will wash the
calcium hydroxide from the surface of the concrete before it has a chance to react to form insoluble calcium
carbonate. It is very rare for secondary efflorescence to
reappear on well-compacted concrete.

3.6.4.3 Minimizing efflorescence
Because many factors influence the formation of efflorescence, it is difficult to predict if and when it will
appear. This fact is evident in the lack of any accepted
standard test method for measuring the efflorescence
potential of concrete. Several experimental methods
have been proposed, but none has been accepted as effectively predicting the performance of concrete in use.
Conditions that increase the penetration of water
into the concrete must be avoided. A dense concrete
that absorbs as little water as possible after curing is
desirable.

ARCHITECTURAL PRECAST CONCRETE

|

253

3

SURFACE AESTHETICS
3.6.4.3 Minimizing efflorescence

In selecting materials for concrete, the soluble-salt
content of all ingredients should be considered. To reduce or eliminate the potential for efflorescence:
1. Reduce concrete permeability and use a concrete with a low water absorption of 5 to 6%
by weight or 12 to 14% by volume. These are
the key factors that the precaster can influence to
minimize efflorescence. Accomplishing these will
require properly graded aggregates, a minimum
cement content for stripping and service strength
requirements (a cement-rich concrete mixture increases water absorption), a low water-cement ratio, good consolidation techniques, thorough curing, and possible use of efflorescence-controlling
agents.
By reducing the amount of available water below
(or to) the amount required for hydration, voids
and capillaries in the matrix are reduced. The reduction of total water content by means of a
water-reducing admixture should reduce the total
porosity slightly, but there are no adequate data
to demonstrate that permeability is reduced materially. However, decreased permeability through
the use of high-range water reducers at equivalent
water-cement ratios has been reported. Reducing
or limiting the water-cement ratio and total water
content to the feasible minimum will greatly aid in
reducing the propensity for efflorescence.
The compound composition of cement of a given
fineness affects permeability of a paste of a given
water-cement ratio at a given age only insofar as
it influences the rate of hydration. The greater the
degree of hydration, the lower the concrete permeability. However, for a given water-cement ratio, coarse cements tended to produce pastes that
are more porous than finer cements.
The permeability of concrete depends on the permeability of the paste as well as that of the aggregate, and on the relative proportions of each.
It also depends greatly on placing, finishing, particularly consolidation, and curing procedures.
Permeability of concrete to liquid water or water
vapor is not a simple function of its porosity, but
depends also on the size, distribution, and continuity of the pores in both the cement paste and
aggregate. The pores in cement paste are of two
kinds. Gel pores, constituting about 28% of the
paste volume, are interstitial spaces between the
gel particles. They are very small (between 1.5 and

254

|

ARCHITECTURAL PRECAST CONCRETE

3.0 nm in diameter). Capillary pores are larger (of
the order of 1 μm) and are irregularly distributed
throughout the cement paste. Because capillary
pores represent the remains of originally waterfilled spaces, their volume can vary between 0 and
40%, depending on the original water-cement ratio and the degree of hydration of the concrete. As
hydration progresses, the permeability decreases.
Thus, normally, the higher the strength of a given
paste or concrete, or the longer it has cured, the
lower it’s permeability.
Air entrainment might be expected to increase
the permeability of concrete. However, because
air entrainment reduces the mixing water requirement and bleeding, and entrained air voids interrupt the continuity of capillary pores, the overall
effect of air entrainment will usually be to reduce
permeability.
2. Use a low-alkali cement. Portland cements vary
appreciably in their total (acid soluble) alkali metal
content (typically 0.02 to 0.90% by weight of cement), dependent on the raw materials used and
the temperature of the kiln. The total alkali content should be limited to less than 0.35% as Na2O
and the water-soluble alkali content to less than
0.13% as Na2O. These severe limitations on alkali
content can be met only by a few cements, other
than portland blast-furnace slag cement. Modern
cement manufacturing methods, which attempt to
conserve energy and reduce kiln emissions, have
led to increased concentrations of alkali metal sulfates, sometimes as much as 1.5% by weight, usually present as soluble sulfates such as K2SO4 or
Ca2K2(SO4)3. It is also suspected that the sulfate content may be as significant as the alkali content in
contributing to efflorescence.
3. Use sand that meets the requirements of ASTM
C 33, C 144, or CSA A 23.1. Never use unwashed
sand containing soluble alkali sulfates. Water-soluble
salts, generally chloride and sulfates, may be deposited in or on sand and gravel deposits by evaporation
of groundwater, or by evaporation of sea or salt lake
water on beaches. Sands may also be contaminated
from soil runoff, plant life, and decomposed organic
compounds. These salt-contaminated materials may
cause efflorescence and should be avoided.
A feature of limestone aggregates is their tendency to exude self-produced efflorescence when
used for exposed-aggregate finishes. On white or

SURFACE AESTHETICS

3.6.4.3 Minimizing efflorescence / 3.6.5 Design of Concrete for Weathering

near-white aggregates this is of little consequence,
and on white finishes it might, on occasion, even
bring some slight improvement to the final result.
On dark surfaces, however, the white film will not
only show, but will significantly dull the original
color of the aggregate. After cleaning the surface
of efflorescence, treatment with a clear sealer can
normally be relied on to prevent a recurrence.

ity, but only marginally in others; their major shortfall is generally the limited amount of time they
provide protection. When considering the use of
an efflorescence-controlling agent, the precaster
should run trial mixtures to determine the effect a
given product may have upon air content and its
compatibility with other admixtures that may be in
the mixture.

4. Use clean mixing water. The mixing water should
be free from harmful amounts of acids, alkalis,
organic material, minerals, and salts. Do not use
drinking water that contains quantities of dissolved
minerals and salts sufficient to cause efflorescence.
Some city drinking-water supplies may require
treatment as they may have as much as 215 ppm
of sodium, 20 ppm of potassium, 550 ppm of bicarbonate, 120 ppm of sulfate, and 280 ppm of
chloride. Do not use seawater.

6. Other factors to consider. Pigments have little
or no effect on efflorescence. Pigments are water
insoluble and do not contain noticeable amounts
of water-soluble salts. Pigments may appear to aggravate any efflorescence problem by making it
more visible. Also, efflorescence deposited on the
surface may mask the true color and give the appearance of pigment fading, even though the pigment itself has not changed.

5. Additives. Additives that reliably prevent efflorescence when added to the concrete are the dream
of every precaster. However, no panacea for the
prevention of efflorescence has been found.
Both fly ash and silica fume consume large quantities of calcium hydroxide (although they are not
generally used in architectural precast concrete because of color problems), as does metakaolin. The
reaction of fly ash with calcium hydroxide is rather slow, so that significant reductions in calcium
hydroxide are seldom seen before 30 to 60 days.
With silica fume additions greater than 20%, it is
possible to almost completely eliminate calcium
hydroxide (although this is not practical). Addition
of 5 to 10% silica fume or metakaolin can produce
large decreases in permeability and alkali content
in the pore structure. Trial mixtures should be used
to evaluate the effect of these materials on air content and other mixture properties and most importantly on color and color uniformity.
Using an integral water-repellent or dampproofer, such as butyl stearate, may reduce the rate of
penetration of water in its liquid state through the
concrete. The butyl stearate is added as an emulsion at a rate of 1% stearate by weight of cement.
Such additives function by blocking capillary action,
providing an internal barrier to the transmission of
water through the matrix. They have performed
very well in some instances, particularly when the
concrete contains paste with relatively high poros-

3

It is important to prevent inadequate hydration of
cementitious materials caused by cold temperatures
or premature drying of the concrete during curing.
Inadequate hydration will prevent the occurrence of
primary efflorescence, but increase the likelihood of
secondary efflorescence.

3.6.5 D
 esign of Concrete for Weathering
The assessment of concrete mixtures with respect to
appearance, strength, and durability is discussed in
Section 3.2.6. This section deals with weathering, not
as it affects structural durability, but with special emphasis on the visual results, particularly staining of the
concrete.
It is apparent that certain concrete surfaces weather
better than others when in the same environment.
Concrete qualities will influence the degree to which
staining of concrete surfaces can be predicted and
limited, and the results of later cleaning of those
surfaces.
The duration of partially wet conditions and the penetration of water, dirt, acidic rainwater, carbon dioxide,
and sulfur dioxide are directly related to the absorption
of the concrete surface. This absorption and penetration also create difficulties in cleaning such surfaces to
restore them to the original appearance.
Low absorption for the surface concrete demands a
high density of concrete and is primarily influenced by
the mixture proportions.
Proportioning concrete for optimum durability and
ARCHITECTURAL PRECAST CONCRETE

|

255

3

SURFACE AESTHETICS

3.6.5 Design of Concrete for Weathering / 3.6.6 Surface Coatings and Sealers

weathering is dependent on the durability of the individual ingredients, as well as the density of the entire
mixture. Concrete should be designed and/or evaluated for each individual project with respect to strength,
absorption, and resistance to freezing and thawing,
where such environments exist.
Optimum quality of concrete for durability and
weather staining should be based on the a low watercement ratio and durable aggregates. Water should
be limited to a minimum as excess water will affect
strength, density, and absorption. Aggregates should
always be checked for potential alkali reactivity. It is recommended that aggregates that are potentially alkali
reactive should not be used for exterior concrete surfaces. This requires a petrographic examination of the
aggregate by qualified personnel according to ASTM
C 295, Standard Practice for Petrographic Examination
of Aggregates for Concrete.
A water absorption test of the proposed facing mixtures may provide an early indication of predictable
weather staining (rather than durability). For the concrete strengths normally specified for architectural precast concrete (5000 psi [34.5MPa] at 28 days), a reasonable water absorption should not be a problem.
Water absorption is an indication of concrete density.
Dense concrete (highly impermeable) is less susceptible to the effects of wetting/drying and, therefore,
will absorb less dirt in a polluted environment. Based
upon the density of normalweight concrete (150 lb/
ft3 [2400 kg/m3]), the water absorption of the proposed face mixture should not exceed 6% by weight.
Alternatively, absorption expressed by volume should
not exceed 14%.
Laboratory freezing and thawing tests have been
conducted to evaluate the durability of concrete under severe climatic conditions. These tests can be made
on prismatic samples prepared from laboratory trial
mixtures or even from cores cut from the face of finished production units. Such tests, however, take several months to complete. The verticality of wall units
seldom allow concrete to reach the critical moisture
saturation point (above 90%) on which such tests are
based. However, where horizontal areas allow water
or snow to accumulate or where ground-level panels
may be subject to splashing by deicing salts and to
freezing conditions at moisture contents above critical saturation, an air-entraining admixture should be
specified. In addition, it is probably a prudent policy

256

|

ARCHITECTURAL PRECAST CONCRETE

to have air-entrained concrete in all precast concrete
exposed to freezing and thawing cycles. Possible exceptions to this are applications that require very high
concrete strengths.
Equivalent evaluations can be obtained more rapidly
by conducting “air void studies” (amount and character of entrained air in cores taken from the production
unit) in accordance with ASTM C 457, Recommended
Practice for Microscopical Determination of Air-Void
Content and Parameters of the Air-Void System in
Hardened Concrete.

3.6.6 Surface Coatings and Sealers
Clear surface coatings or sealers may be considered
for the possible improvement of concrete’s weathering
characteristics. The quality of concrete normally specified for architectural precast concrete, even with the
minimum practical thickness, is waterproof.
Sealers may be applied for the following reasons:
1. The prime justification for their use is the potential
improvement of weathering qualities in urban or industrial areas. A sealer may be used to reduce attack
of the concrete surface by airborne pollutants.
2. To facilitate cleaning of the surface if it becomes
soiled or to resist impregnation of graffiti.
3. To prevent changes in appearance, particularly darkening of surfaces that are wetted, by making them
water repellent. Also, uneven drying is caused by
untreated concrete absorbing moisture at different
rates over its surface. The surface will then dry unevenly, leaving a patchy appearance. Sealers keep
the moisture at the surface. Drying is sped up and
the surface retains its natural appearance.
4. To reduce efflorescence, particularly with a gray or
dark cement matrix. The use of a concrete sealer
will reduce the absorption of moisture into the surface, thereby minimizing or eliminating the wetdry cycle and the subsequent migration of water
and salts to the surface.
5. To reduce the incidence of concrete surface leaching, which may be a factor in the etching of the
glass, aluminum, and other lime-susceptible construction materials. However, a sealer is not a substitute for proper design for water run-off.
6. To reduce the tendency of soiling in the yard, in
transportation, and on the building. Special sealers
(often with short-term effect) that will not change

SURFACE AESTHETICS

3.6.6 Surface Coatings and Sealers

the appearance of the units are used by some precasters to protect the units during yard storage,
particularly where dirty atmospheric conditions
exist.
7. To brighten aggregates and develop color tones
that would otherwise be subdued.
The effectiveness of sealers in any of the preceding
applications is dependent on the properties of the specific sealers. Some problems that have occurred with
certain sealers are:
1. Appearance changes vary with the age of the precast concrete unit being treated. If a methyl methacrylate resin sealer with a high solids content is
applied, the unit will take on a glossy or wet look.
Some sealers may create a blotchy appearance if
applied before the units are fully cured and dried,
while others may lose effectiveness if sealers are
applied too soon after unit fabrication. Sealers
may also accentuate a patched area of different
density. If it is necessary to remove sealers, use a
solvent or wire brush, or grind or lightly sandblast
the surfaces.
2. Certain sealers, such as some silicones, have been
found to attract airborne hydrocarbons.
3. Sealers will often interfere with patching of the
precast concrete surface or with adhesion of joint
sealants. Some methyl methacrylate resin sealers
inadvertently sprayed in the joints may peel away
from the concrete surface leaving a void between
sealant and concrete, while silicone water repellents in the joints may prevent adhesion of joint
sealant to the concrete surface. Therefore, the
sealer/sealant compatibility should be verified.
Application of sealers should, therefore, be delayed
until all repairs, cleaning, and sealant applications
are completed.
4. The possibility of severe and permanent discoloration of the concrete surface to various shades of
yellow, brown, or gray, or of peeling varies considerably between types and sources of sealers.
Moisture permeability (breathing) of the sealer is
a requirement to prevent blistering and peeling of
the sealers. The sealer must make the concrete surface less water absorbent while allowing the outward transmission of water vapor, permitting the
surface to breathe.
5. Proper application by following the sealer manu-

3

facturer’s instructions depends on qualified operators and possible expensive pretreatment of the
precast concrete units. The application limitations
of sealers with respect to timing, ambient temperatures, moisture content of the concrete, and
method and rate of application should be fully
investigated before choosing a particular type of
sealer or supplier.
6. Uncertain life expectancy, possibly causing a maintenance expense through resealing.
The use of sealers on precast concrete in locations
having little or no air pollution or in dry climates is not
recommended due to the additional cost, recurring periodic maintenance applications, and uncertain results
of the sealer application.
A careful evaluation should be made before deciding
on the type of sealer. This should include consultation
with local precasters. Suggested sealers should be tested on reasonably sized samples with varying ages. The
performance and affect on overall wall appearance
should be monitored over a suitable period of time or
be based on prior experience under similar exposure
conditions. Sealers should be guaranteed by the supplier or applicator not to stain, soil, darken, or discolor
the precast concrete finish. Also, some clear coatings
(sealers) may cause sealants to stain the concrete or
affect the bond of the joint sealant. The manufacturers
of both joint sealant and sealer should be consulted
regarding compatibility prior to application or the materials should be tested before application.
The type of solvent used in sealers, as well as the
solids content, can affect the resulting color of the
concrete surface. Thus, neither the type nor source of
sealer should be changed during the project. Generally,
sealers having higher solids contents tend to produce
darker surfaces with glossy effects. The amount of color change depends primarily on the types of material
in the sealer and their concentration, as well as the
porosity of the concrete surface. The active ingredient
of the sealer must be chemical-resistant to the alkaline
environment of the concrete. Also, the sealer should
dry to a tack-free finish to prevent a build-up of airborne contaminants that results in surface staining.
The sealers should be evaluated on how well they
penetrate concrete surfaces that vary in absorption
and texture. The penetrating sealers, generally silanes
or siloxanes, develop their water repellent ability by
penetrating the surface to depths of up to 1/4 in. (6

ARCHITECTURAL PRECAST CONCRETE

|

257

3

SURFACE AESTHETICS
3.6.6 Surface Coatings and Sealers

mm), reacting with the cementitious materials in the
concrete and making the concrete hydrophobic, but
they do not have crack bridging capabilities. Some water repellents seal cracks by rendering the crack sides
water repellent. The penetrating ability of the sealer
system depends on the molecular size of the active ingredient, the viscosity of the system, and the solventcarrying system.
Silane products based on monomeric alkylalkoxysiloxane (AS) have an extremely small molecule that provides excellent penetrating power. However, while the
active substance is being formed, the relatively volatile
silane can evaporate. The rate of evaporation increases
as the drying conditions increase and as high concentrations of active ingredient are used (typically 20 to
40% silane).
Silane products based on oligomeric alkylalkoxysiloxane (OAS) also have an extremely small molecular
structure, but they have practically no vapor pressure
under application conditions. This means that they do
not evaporate readily and can remain in capillaries until
conditions are favorable. They can, therefore, be used
at much lower levels (5 to 10%) of active ingredients.
While generally more expensive than other types of
sealers, silane sealers are recommended and are typically longer lasting and less subject to deterioration
from ultraviolet light exposure. Because appearance or
characteristics of the surface is generally unaffected by
application of these sealers, it is usually difficult to detect where they have been applied.
Surface sealers consisting mainly of the methyl methacrylate form of acrylic resins, having a low viscosity
and high solids content also produce durable finishes
but usually result in glossy surfaces. A combination of
a base coat application of a penetrating sealer with
a topcoat application of a methacrylate-based sealer
may be the most effective.
In most cases, except for the silane or siloxane sealers,
it may be necessary to adequately roughen an as-cast
surface in order to obtain good adhesion of the coating.
One way is to lightly sandblast the surface to be coated.
Surface coatings should not be applied until joints are
caulked and all repairs and cleaning have been completed. In cases where the precast concrete units have
been coated at the manufacturing plant and additional cleaning is required, it may be necessary to recoat
those particular units. If panels are only recoated in

258

|

ARCHITECTURAL PRECAST CONCRETE

spots, this could lead to inconsistencies in color.
Sealers should be applied in accordance with the
manufacturer’s written recommendations. Silanes and
siloxanes are best applied to slightly damp surfaces.
Generally, low-pressure, 15 to 30 psi (0.1 to 0.2 MPa),
airless spray equipment should be used to apply the
sealer. This results in a uniform application while avoiding excessive sealer rundown. Two coats are usually required to provide a uniform coating, because the first
coating is absorbed into the concrete. The second coat
does not penetrate as much and provides a more uniform surface color. Care should be taken to keep the
sealer off glass or metal surfaces, unless testing shows
no detrimental effect.
Graffiti repellents: Anti-graffiti coatings are either
permanent or sacrificial and should not be confused
with water repellent materials mentioned previously.
Successful coatings tend to be slick or shiny. They offer no “tooth” for graffiti materials to cling. The coatings also tend to retard the wall’s ability to breathe.
Therefore, their use should be limited to those areas
subject to graffiti—generally within about 8 ft (2.4 m)
of grade. These coatings may change the color of the
precast concrete by altering the refractive qualities of
its surface; thus, they may become a design consideration. ASTM D 6578 Practice for Determination of
Graffiti Resistance provides a protocol for evaluating
coatings for graffiti resistance. The tests are performed
on smooth surfaces. While the coating may have a
certain cleanability level, the roughness of the surface
would affect the ability to adequately rub the graffiti
completely over the surface. The graffiti in the texture
created by the substrate may be difficult to remove unless the proper tools and procedures are used.
Aliphatic urethanes are considered the best anti-graffiti coatings because of their resistance to solvents, yellowing, and abrasion. Solvents such as mineral spirits or
methyl ethyl ketone can remove most graffiti from an
aliphatic polyurethane without compromising the urethane coating. Acrylics, epoxies, silanes, and siloxanes
are also used to make graffiti removal easier; however,
acrylics dissolve with the solvent and epoxies tend to
yellow or discolor. Silanes and siloxanes may not resist
certain graffiti materials as well as the urethanes, but
they do maintain a high breathability at the surface
while resisting penetration of graffiti materials into the
pores of the concrete.

SURFACE AESTHETICS

3.6.7 Maintenance and Cleaning

3.6.7 Maintenance and Cleaning
Precast concrete units require little maintenance to
preserve their original appearance. By following a simple program of inspection and maintenance, precast
concrete can easily achieve the design service life of a
building. To ensure proper performance or appearance,
it is recommended that visual inspections be carried
out yearly. Attention should be given to the caulked
joints, surface appearance, and connections, if visible.
Minor problems discovered and addressed in a timely
manner will prevent expensive future repairs.
There are specific items that require periodic attention:
1. Window cleaning—clean every 90 to 180 days
(based on dirt accumulation effects of the environmental pollution).
2. Dirt removal—the precast concrete façade may be
power-washed as necessary (based on the effects
of the environmental pollution). Buildup of dirt is
usually a gradual process, and a periodic flushing
with plain water may be an adequate maintenance
program.
3. Joint sealants—check all joint sealants for deterioration and repair, as required. Typical maintenance
issues are water leakage through the joint, visible
separation of the sealant from the concrete, and
cracking or tearing of the sealant. Sealants generally require re-caulking only every 15 to 20 years.
	Periodic evaluation of a building façade can help
detect sealant deterioration before joints have
failed and let water into the building. Building
owners should keep accurate records of when
their exterior sealants were installed and the average useful service life of the sealant. Once the sealant has reached 75 percent of its useful life, periodic review of the sealants should be conducted.
In most cases, the initial evaluation could be done
from the ground and the roof. Once the sealants
have reached the average useful service life, a more
extensive evaluation should be performed, including the use of a swing stage to adequately observe
the building sealant joints.
4. Sealer—if methyl methacrylate sealer was applied,
it should be re-applied every 4 to 5 years or as
specified by the manufacturer; if penetrating silane
or siloxane sealer was applied, it may not be necessary to recoat. If desired to recoat, the minimum
time would be 7 to 10 years.

3

Precautions should be taken to avoid damaging or
staining precast concrete units by:
1. Ensuring access equipment does not scratch or
chip precast concrete surfaces.
2. Ensuring window cleaning solution (run-off) is
cleaned from precast concrete units to prevent
staining.
Removing stains from old concrete sometimes leaves
the area much lighter in color than the surrounding
concrete because surface dirt has been removed along
with the stain or because the surface may have become
slightly bleached. If at all possible, cleaning of the precast concrete should be done when the temperature
and humidity allow rapid drying. Slow drying increases
the possibility of efflorescence and discoloration. There
is no single prescription for the cleaning restoration of
architectural precast concrete as each building is exposed to a unique set of ambient conditions.
Because efflorescence often occurs during or immediately following construction, the first impulse is to immediately wash it off with water or an acid cleaning solution. This is not advisable, particularly in cool or damp
weather when the primary result of such action will be
to introduce more water into the concrete. The water
will wash some of the alkali salts from the surface but
will also dissolve and carry the salts back into the concrete, thus causing a reoccurrence of the efflorescence.
If it is possible to wait for one to two years before doing anything to the building, 95% of the time, the efflorescing salts will work themselves to the surface; the
problem may solve itself by normal weathering. The
water-soluble alkali salts will gradually weather away.
Heavy calcium carbonate efflorescence, although less
common, is extremely difficult to remove as it forms
a hard, white crust. After weathering to calcium hydrogen carbonate, it may be easily removed, otherwise
acidic cleaners may be necessary.
It is often helpful to determine the type of efflorescent salt, dirt, or stain so that a cleaning solution can
be found that readily dissolves it without adversely affecting the surface finish.
Before cleaning precast concrete, a small (at least [3
x 3 ft (0.9 x 0.9 m)]), inconspicuous area should be
cleaned and checked to be certain there are no adverse
effects on the concrete surface finish or adjacent corrodible materials such as glass, metal, or wood, before
proceeding with the cleaning. A sprayed-on, strip-off

ARCHITECTURAL PRECAST CONCRETE

|

259

3

SURFACE AESTHETICS
3.6.7 Maintenance and Cleaning

masking can be used to protect glass and aluminum
frames. The effectiveness of the cleaning method on
the sample area should not be judged until the surface
has dried for at least one week.
The key to successful cleaning is recognizing the advantages and limitations of each technique and designing a cleaning program around them (see ASTM E
1857, Guide for Selection of Cleaning Techniques for
Masonry, Concrete, and Stucco Surfaces). A suggested
order for testing appropriate cleaning procedures for
removal of dirt, stains, and efflorescence from precast
concrete (beginning with the least damaging) is:
1. D
 ry scrubbing with a stiff nylon fiber brush, particularly if the surface is brushed shortly after the
appearance of dirt or efflorescence.
2. Wetting the surface with water and vigorous scrubbing of the finish with a stiff fiber brush followed
by thorough rinsing of the surface with clean
water. Low-pressure water, 50 to 200 psi (0.3 to
1.4 MPa); spraying (water misting); high-pressure
water, 400 to 800 psi (2.8 to 5.5 MPa); or steam
cleaning, 10 to 80 psi (0.07 to 0.55 MPa) may also
be tried to remove dirt. Steam cleaning is also done
in conjunction with chemical cleaning. Disposal of
run-off water from washing needs to consider environmental compliance requirements.
3. Chemical cleaning compounds such as detergents,
muriatic or phosphoric acid, or other commercial
cleaners used in accordance with the manufacturer’s recommendation. If possible, a technical representative of the product manufacturer should be
present for the initial test application to ensure its
proper use. Consideration should be given to the
chemical’s effect on the concrete surface finish and
adjacent materials.
Areas to be cleaned chemically should be thoroughly saturated with clean water prior to application of the cleaning material to prevent the chemicals from being absorbed deeply into the surface
of the concrete. Surfaces should also be thoroughly rinsed with clean water after application so that
no trace of chemicals remain in the surface layers of the concrete. Cleaning solutions should not
be allowed to dry on the concrete finish. Residual
salts can flake or spall the surface or leave difficult
stains. Misapplication of hydrochloric acid can lead
to corrosion of adjacent or embedded metals that
have shallow cover. Care should be taken to pro-

260

|

ARCHITECTURAL PRECAST CONCRETE

tect all corrodible materials, glass, or exposed parts
of the building during acid washing.
Care should be taken to use dilute solutions of
acid to prevent surface etching that may reveal the
aggregate and slightly change surface color and
texture of the precast concrete, affecting the appearance of the finish. The entire precast concrete
façade should be treated to avoid discoloration or
a mottled effect. Application should be to small
areas of not more than 4 ft2 (0.4 m2) at a time, with
a delay of about five minutes before scouring off
the deposit with a stiff bristle brush. Any of several
diluted solutions of acids are effective ways to remove dirt, stains, and efflorescence.
a. 1 part hydrochloric acid in 9 to 19 parts water
b. 1 part phosphoric acid in 9 parts water
c. 1 part phosphoric acid plus 1 part acetic acid
(vinegar) in 19 parts water
d. 1 part acetic acid in 5 parts water
Hydrochloric (muriatic) acid may leave a yellow
stain on white concrete. Therefore phosphoric or acetic acid should be used to clean white
concrete.
Rubber gloves, glasses, and other protective clothing must be worn by workers using acid solutions
or strong detergents. Materials used for chemical
cleaning can be highly corrosive and frequently
toxic. All precautions on labels should be observed
because these cleaning agents can affect eyes,
skin, and breathing. Materials that can produce
noxious or flammable fumes should not be used
in confined spaces unless adequate ventilation can
be provided.
4. Dry or wet abrasive blasting using sand or other
abrasives may be considered if this method was
originally used in exposing the surface of the unit.
Excessive abrasive blasting may change the color
and texture of the finished unit and must be avoided. An experienced subcontractor or a precaster
should be engaged for sandblasting. Abrasive
blasting with industrial baking soda will not affect
the concrete surface. (Any residue on the surface
must not be removed by water as efflorescent salts
may be dissolved and carried into concrete causing additional efflorescence.) Residues should be
blown, vacuumed, or brushed from the surface.
Stone veneer–faced precast concrete units should be

SURFACE AESTHETICS

3.6.7 Maintenance and Cleaning

cleaned with stiff bristle, stainless steel, or bronze wire
brushes, a mild soap powder or detergent, and clean
water using low or high pressure depending on stone
type, if necessary. Acid or other strong chemicals that
might damage or stain the stone veneer should not be
used. Information should be obtained from stone suppliers on methods of removing oil, rust, and dirt stains
from the stone.
Mortar stains may be removed from brick-faced panels by thoroughly wetting the panel and scrubbing
with a stiff bristle brush and a masonry cleaning solution. A prepared cleaning compound is recommended;
however, on red brick, a weak solution of muriatic acid
and water (not to exceed a 10% muriatic acid solution)
may be used. Acid should be flushed off the panel with
large amounts of clean water (using a pressure washer)
within 5 to 10 minutes of application. Brick should be
cleaned in accordance with the brick manufacturer’s
recommendations, possibly using proprietary cleaners
rather than acid to prevent green or yellow vanadium

3

stains and brown manganese stains.
Following the application of the cleaning solution,
the panel should be rinsed thoroughly with clean water. Low pressure using a 30 to 50 psi (0.2 to 0.3 MPa)
washer or high-pressure water cleaning techniques
may also be used to remove mortar stains except on
sand finished brick.
Unglazed tile or terra cotta surfaces should be cleaned
with a 5% solution of sulfamic acid for gray or white
joints, and a more dilute (2%) solution for colored
joints. The surface should be thoroughly rinsed with
clean water both before and after cleaning. Glazed tile
manufacturers generally do not recommend the use of
acid or abrasive powders for cleaning purposes.
For information on removing specific stains from concrete, reference should be made to Removing Stains
And Cleaning Concrete Surfaces, IS 214, published by
the Portland Cement Association, Skokie, IL.

ARCHITECTURAL PRECAST CONCRETE

|

261

4

Park Tower,
Chicago, Illinois;
Architect: Lucien Lagrange & Associates Ltd., Chicago; HKS Inc., Dallas.

CHAPTER FOUR
DESIGN
4.1 D
 ESIGN AND CONSTRUCTION
RESPONSIBILITY
4.1.1 G
 eneral
Design and construction of a structure is a complex
process. Clearly defining the scope of work and the
responsibilities of the involved parties by means of the
contract documents is critical to achieving the desired
result. This section provides a guide for all parties involved in a precast concrete project and defines the
responsibilities of each party. These responsibilities and
relationships between the parties are defined in the
contract documents for a particular project.
A successful precast concrete project requires teamwork, close cooperation and coordination between all
of the participants, including the owner, architect, engineer of record (EOR), precast concrete manufacturer,
erector, general contractor (GC)/construction manager
(CM), and all other affected trades. The scope of the
precast concrete work and the responsibilities of each
party (typically defined by the contract documents)
should be established at an early stage in the development of a project to achieve the desired quality and
keep the project on schedule (see Table 4.1.1). During
construction, each party is responsible for communicating with all other parties through the GC/CM or architect. This helps to prevent misunderstandings and
confusion. When authority and responsibility roles are
clearly defined by the contract documents, problems
and conflicts are avoided. Local practices regarding the
assignment and acceptance of responsibility in design
and construction can vary.
One of the basic principles of the construction industry is that responsibility and authority should go hand
in hand. Another principle is that every party should
be responsible for its own work. These principles are
frequently not followed in practice. There have been
cases where owners have sued architects or engineers
for approving non-conforming work without giving
them authority to monitor the work as it progressed.
Safety enforcement agencies (OSHA) and plaintiffs’
lawyers have charged engineers or architects with the
responsibility for construction accidents contrary to
language and responsibilities listed in the contract doc-

uments. These last two situations typically are cases of
responsibility without authority, although there could
be instances where a design team’s work or direction
can affect jobsite safety. If the design team is involved
with construction-management functions, they could
be making decisions affecting worker safety as well
as quality of construction. When agents of the owner
give instructions directly to the construction workforce
regarding how work is to be performed, they step over
the line into the contractor’s area of responsibility.
The increased complexity of structures today makes it
essential to have design input from the subcontractors.
This input, whether submitted as value engineering
proposals, in response to performance requirements, or
simply offered as design alternatives, plays a legitimate
role in construction. For example, a precast concrete
subcontractor may propose alternatives that improve
the efficiency of the fabrication or erection operation.
In approving the alternatives, the design team retains
responsibility for properly interfacing with other materials in contact with or adjacent to the precast concrete
elements.
The EOR always has to take overall responsibility for the
structural design of the completed structure. However,
certain aspects of the design are often delegated to
specialty engineers working for the material suppliers
or subcontractors. When any of this delegated structural design work for a portion of the structure involves
engineering, the design work should be reviewed and
approved by the EOR registered in the same state as
the project or as required by the local jurisdiction. The
EOR then accepts responsibility for the overall structural
design. Additionally, local regulatory authorities should
be consulted for their specific requirements. Contract
documents typically require the structural design be the
responsibility of a professional engineer, regardless of
conflicts with other governmental requirements.

4.1.2 Responsibilities of the Architect
The architect develops the project design concept,
establishes overall structure geometry, selects the cladding material for appearance and function, provides
details and tolerances for proper material interfacing

ARCHITECTURAL PRECAST CONCRETE

|

263

4

DESIGN

4.1.2 Responsibilities of the Architect

Table 4.1.1. Design Responsibilities

Contract Information Supplied
by Design Team

Responsibility of the Precaster
OPTION I

Provide complete drawings
and specifications detailing all
aesthetic, functional, and structural
requirements including design
criteria, plus dimensions.

The precaster should make shop drawings (erection and production
drawings), as required, with details as shown by the designer.
Modifications may be suggested that, in precaster’s estimation, would
improve the economics, structural soundness, or performance of the
precast concrete installation. The precaster should obtain specific approval
for such modifications. Full responsibility for the precast concrete design,
including such modifications, remains with the designer. Alternative
proposals from a precaster should match the required quality and remain
within the parameters established for the project. It is particularly advisable
to give favorable consideration to such proposals if the modifications
are suggested so as to conform to the precaster’s normal and proven
procedures.
OPTION II

Detail all aesthetic and functional
requirements but specify only the
required structural performance
of the precast concrete units.
Specified performance should
include all limiting combinations of
loads together with their points of
application. This information should
be supplied in such a way that all
details of the unit can be designed
without reference to the behavior
of other parts of the structure. The
division of responsibility for the
design should be clearly stated in
the contract documents.

The precaster has two alternatives:
(a) S ubmit erection and shape drawings with all necessary details and
design information for the approval and ultimate responsibility of the
designer.
(b) S ubmit erection and shape drawings, and design information for
approval and assume responsibility for the panel structural design; that
is, the individual units, but not their effect on the building. Precasters
accepting this practice may either stamp (seal) drawings themselves,
or commission engineering firms to perform the design and stamp the
drawings.
The choice between the alternatives (a) and (b) should be decided between
the designer and the precaster prior to bidding with either approach clearly
stated in the specifications for proper allocation of design responsibility.
Experience has shown that divided design responsibility can create contractual problems. It is essential that the allocation of design responsibility is
understood and clearly e×pressed in the contract documents.
OPTION III

Cover general aesthetic and
performance requirements only and
provide sufficient detail to define
the scope of the precast concrete
work.

The precaster should participate in the preliminary design stage and the
development of the final details and specifications for the precast concrete
units and should work with the design team to provide an efficient design.
The precaster provides the engineering design of the precast concrete
units and their connections to the structure and should work with the
design team to coordinate the interfacing work. The precaster should
submit design information for approval and shop drawings at various
stages of completion for coordination with other work.

and weatherproofing, and specifies performance characteristics, as well as inspection parameters and testing
requirements in the contract documents.
The architect and EOR have responsibility to coordinate the design aspects of the precast concrete panels

264

|

ARCHITECTURAL PRECAST CONCRETE

such as aesthetics, dimensions and loads to structure.
The architect or EOR may specify in the contract documents that design services for portions of the work
are to be provided by the precaster. Typically design
services are performed for the precaster by a licensed
engineer who can be an employee of the precaster or

DESIGN

4.1.2 Responsibilities of the Architect

an independent structural engineer. The contract documents should clearly define the scope of the precast
concrete design requirements and document review
responsibilities, as well as the responsibilities of otherparties providing design services.
The contract drawings prepared by the design team
should provide the overall geometry and dimensions of
the structure, member or panel dimensions and crosssections, typical connection locations and details, and
concepts so all precasters are estimating based on the
same information. The architect’s drawings may only
show reveals or design articulation, allowing the precaster to determine panel sizes suitable to their handling and erection capabilities. In addition, the contract
documents (specifications and design drawings) also
should provide the general performance criteria, design loads, including concrete strength requirements,
deflection requirements, temperature considerations,
and any tolerance or clearance requirements for proper interfacing with other elements of the structure.

4

concrete engineer of their design responsibilities.
Key design issues for the design team. The contract drawings prepared by the design team should
provide a clear representation of the configurations
and dimensions of individual precast concrete units
and their relationship to the structure and to other materials. Contract documents that are unclear and lack
detail may extend shop drawing preparation time, lead
to confusion over work scope, and impact the project
schedule.
The contract documents should supply the following
information:
• Elevations, sections, and dimensions necessary to
define the sizes and shapes (profiles) of each different type of precast concrete element;
• Locations of joints, real (functional) or false (aesthetic);
• Required materials, color and finish treatment for
all surfaces with a clear indication of which surfaces
are to be exposed to view when installed;

The order in which the project contract, specifications,
or drawings prevail in the event of conflicts should be
clearly defined. All aesthetic, functional, and structural
requirements should be detailed.

• Corner details;

The design team should provide complete, clear,
and concise drawings and specifications. Contract
documents should clearly define: (1) precast concrete
components that are to be designed by the precaster
(state who takes responsibility for design of elements
at interfaces with other parts of the structure, such as
the secondary steel bracing of the structure, to prevent
rotation of beams or columns); (2) details or concepts
of supports, connections, and clearances that are part
of the structure designed by the design team and that
will interface with the precast concrete components;
and (3) permissible design load transfer points and indicate generic connection types to avoid having the
precaster make assumptions on connection types and
piece counts during bidding and design. It is preferable to leave specific panel and connection design to
precasters so they can design details and connections
suitable for their production and erection techniques.

• Openings for services and equipment, with their approximate size and location;

The architect and EOR should review designs, calculations, and shop drawings submitted by the precaster
for conformance with design criteria, loading requirements, connection points, and design concepts as
specified in the contract documents. This review, however, does not relieve the precaster and the precast

The precaster uses the information from the contract
drawings and documents to generate shape and erection drawings and design calculations. These drawings
should detail elevations showing panel sizes, surface
features, and panel relationships; detail sheets should
show panel cross-sections, special edge conditions,

• Details for jointing and interfacing with other materials (coordinated with the general contractor), including windows, roofing, and other wall systems;

• Details for special or unusual conditions including
fire endurance requirements;
• Governing building codes, design loads, and deflection limitations;
• Specified dimensional tolerances for the precast
concrete and the supporting structure, location tolerances for the contractors’ hardware, clearance requirements, and erection tolerances for the precast
concrete. Exceptions to PCI MNL-117 or MNL-135
tolerances are not recommended;
• Support locations for gravity and lateral loads;
• Building location and site access; and
• Delineation of lateral bracing for structural beams
or any unusual erection sequence requirements.

ARCHITECTURAL PRECAST CONCRETE

|

265

4

DESIGN

4.1.2 Responsibilities of the Architect / 4.1.3 Responsibilities of the Engineer of Record

and feature details; and should specify connection details showing mechanisms and locations of load transfers to the supporting structure. Allowing the precaster
to suggest configurations of the precast concrete units
and the opportunity to select which joints are false and
which are real (panelization) will achieve greater economy and flexibility in production and erection.
The design team should review shop drawings in a
timely manner to ensure their general conformance
with the contract documents, to avoid delay in the
project schedule, and to respond to aesthetic questions raised by the construction team. Architectural
and structural review and clarification of dimensions
and detailing should be anticipated. Following this review, the precaster will make the appropriate revisions
to the shop drawings. Open discussion between the
architect and precaster should be allowed and encouraged in order to achieve the best possible design for
the project.
Producing small mockups is encouraged to help verify
the appearance of the completed façade and clarify
actual field-construction techniques and material interface issues. If the units have returns, the same size
return should appear in the mockup panels.
The architect establishes the standards of acceptability for surface finish, color range, and remedial procedures for production and construction defects and
damage. This can be best accomplished by the precaster producing at least three sample panels, 15 to 20 ft2
(1.4 to 1.9 m2) each, before the initial production to establish the range of acceptability with respect to color
and texture variations, surface blemishes, and overall
appearance. In addition the architect should visit the
plant during the first week of production to evaluate
conformance with approved samples.
Panel-to-panel joint design and the proper sealing at
windows and other penetrations in the exterior wall
is necessary to prevent air and water infiltration. The
architect is responsible for providing these designs and
details. Precast concrete is inherently watertight and
impermeable and therefore it is important to have watertight joints at the window-to-precast concrete interface to prevent water leaks. The architect should examine and modify these details, as required. The contract
documents should require that the same sealant contractor seal all joints in order to avoid sealant incompatibility thereby providing single source responsibility.

266

|

ARCHITECTURAL PRECAST CONCRETE

For large projects or for special conditions where moisture protection is a concern, specifications can call for the
production, shipping, and erection of a full-scale mockup at a testing lab. This mockup would include various
precast concrete and window elements assembled and
caulked. While, a wind-driven rain test, can be costly and
time consuming, it can verify moisture protection details
and satisfy any moisture penetration concerns or requirements. The cost of these tests must be included in the
project budget. These mockups and tests can be expensive and should be specified only where there is a demonstrated need. When such tests are needed, sufficient time
must be provided in the project schedule to evaluate the
test results and incorporate any consequent modifications
into the final design.
After the product is erected and detailed, the architect
should promptly prepare a punch list setting forth, in accurate detail, any items of the work that are not found
to be in accordance with the contract documents so
that proper corrective action may be taken. A meeting
between the contractor, precaster, erector, and design
team should then be held promptly to discuss any questions concerning what the design team requires to be
done before the work can be accepted as complete. All
repairs should conform to the contract documents and
the architect’s requirements (for matching the color and
finish of the approved sample) and should be structurally
sound. If the repairs cannot be completed to a satisfactory level the repairs may be rejected. The industry standard for evaluating the visual acceptability of repairs is at
a 20 ft (6 m) viewing distance with the unaided eye.
When advised by the precaster that the punch list items
have been completed, the GC/CM and design team
should check the corrections. After the precast concrete
units have been accepted, subsequent responsibility and
liability for their condition rest with the GC/CM.

4.1.3 Responsibilities of the Engineer
of Record
The EOR has responsibility for specifying the design
criteria for the design of the precast concrete elements
and for describing the intended load paths. The EOR
should anticipate the loadings in the structural design
and provide a structural system adequate to support
these loads. The EOR should define the type of loading to be applied to the panels and the structure, as
well as provide information, applicable codes (design
criteria), including wind, seismic or blast design, when

DESIGN

4.1.3 Responsibilities of the Engineer of Record / 4.1.4 Responsibilities of the General Contractor/Construction Manager

applicable. The EOR should consider the consequences
of the eccentricities of the weight of the precast concrete panels when designing the supporting structure.
Any special erection procedures or sequences should
be clearly defined, prior to bidding, in the contract
documents. For example, can one elevation be erected
at a time (less crane movement), or must the erection
be one level at a time to prevent undue stresses on
the structure? Observations in the field have shown
that where precast concrete panels are erected to a
greater height on one side of a multistory building
than on the other, the steel framing can be pulled out
of alignment. Precast concrete panels can be erected
at a relatively uniform rate around the perimeter of the
structure or the designer of the structural frame should
determine the degree of imbalanced loading permitted. Other limitations may involve the rigidity of the
structure, requiring that walls not be erected prior to
completion of floors designed to carry the lateral loads.
The EOR has the responsibility of reviewing the precast
concrete design work for compatibility with the overall
structural design and structual stability. This does not,
however, relieve the EOR from the overall design responsibility for the safety and proper performance of
the completed structure.
The EOR should determine and show on the contract documents the locations for supporting the gravity and lateral loads of the precast concrete units, including intermediate lateral (tieback) connections, if
necessary. The EOR’s review of the erection drawings
confirms that the structure is adequate, within defined
deflection limitations, to resist the anticipated loads
and forces from the precast concrete, and verifies that
the magnitude and location of the loading points on
the structure agree with the original design intent. It
is important that preliminary meeting(s) between the
architect, EOR, and precaster be held before structural
members are ordered and fabricated so panel sizes,
shapes, and basic connections and their locations can
be established. For steel frame structures, the EOR
should determine how far in advance the final connections of the frame must be completed prior to precast
concrete panel erection.
The gravity supports of precast concrete panels are
generally eccentric to the centerline of the supporting steel or concrete members. The EOR should design
the structural members to prevent excessive deflection
and rotation of the supporting structure during and
after erection of the precast concrete, as well as de-

4

termining the need for diagonal bracing or stiffening
of supporting structural members. Supplemental framing necessary to support the precast concrete should
be noted on the structural drawings. Responsibility for
designing, supplying, and installing the bracing for the
structure and the secondary steel should be clearly addressed in the contract documents and discussed in a
prebid meeting. Typically, the steel subcontractor supplies all supplemental support, such as diagonal bracing and stiffeners based on the EOR’s design, and coordinates locations with the precast concrete erection
drawings.

4.1.4 Responsibilities of the General
Contractor/Construction Manager
The responsibilities of the CM, who is engaged by
the owner to manage and administer the construction,
may be different from those of the GC, depending on
the CM’s agreement with the owner and local practice.
The responsibilities of the CM, while generally similar
to those of the GC, should be clearly defined in the
contract documents.
The GC/CM should have the responsibility and authority of implementing the design intent of the contract documents, which includes furnishing materials,
equipment, and labor; maintaining specified quality
and schedule requirements; and coordinating of all
trades. The GC is responsible for construction means,
methods, techniques, sequences, and procedures.
Also, the GC should initiate, maintain, and supervise
all safety procedures and programs on the construction site. Site access to the structure for erection of
the precast concrete elements is an important issue.
The GC is responsible for providing and maintaining
clear, level, well-drained unloading areas and stabilized
road access around and into the structure so the hauling and erection equipment are able to operate under
their own power.
The GC/CM generally has no direct design responsibility but does, however, have considerable impact on the
design process through their coordination role. The GC/
CM is responsible for coordinating the information necessary to allow the preparation of the precast concrete
erection drawings as well as reviewing and securing approval for the shop drawings, samples, mockups, and
range samples. The GC/CM receives the shop drawing
submittals from the various trades and together they
form the completed project design. The GC/CM is re-

ARCHITECTURAL PRECAST CONCRETE

|

267

4

DESIGN

4.1.4 Responsibilities of the General Contractor/Construction Manager

sponsible for the timely transmission and resolution of
requests for information (RFI). The GC/CM is normally
responsible for project schedule, grid dimensions at
each floor level (which includes control points, benchmarks, lines on the building, and work points for angled
or curved building elevations), so all trades are working from uniform data and common reference points.
Dimensional interfacing of the precast concrete with
other materials and construction trades, and the maintenance of the structure’s specified tolerances to ensure
proper fit, is also a responsibility of the GC/CM. The GC
should notify the precaster and erector when as-built
conditions (dimensions) of the structural framing vary
beyond the tolerances stated on the contract drawings.
Dimensional tolerances between interfacing materials,
such as precast concrete units and glazing, should also
be considered.
The GC/CM should encourage direct communication between the precaster, EOR, and the architect.
All communications should be confirmed in writing and distributed to all parties in order to avoid
misunderstandings.
Typically, the GC is responsible for placing embedded items in cast-in-place concrete and coordinating
steel attachments with the steel fabricator according
to a layout or anchor plan supplied by the precaster.
In most instances, the most economical approach is to
have required connection hardware attached to steel
columns or beams by the steel fabricator. This necessitates awarding the precast concrete contract simultaneously with the steel contract so that early coordination between these trades can occur. Changes to
panel bearing surface and anchorage locations other
than adjustments within prescribed tolerances require
approval by the design team. The GC/CM should provide the precaster with as-built surveys of embedded
items, anchor bolts, and other attached hardware so
that misaligned or missing hardware can be identified
and remedial actions undertaken by GC/CM prior to
erection of precast concrete units.
For concrete frames, the GC/CM should provide the
erector with authorization to begin erection after the
concrete has reached design strength and any interfering
formwork or shoring has been removed. For steel frame
structures, the GC/CM should provide the erector with
the authorization to begin erection after the steel frame
has been adequately detailed and stabilized, which is
typically after concrete floors have been placed.

268

|

ARCHITECTURAL PRECAST CONCRETE

After erection of the precast concrete panels, the GC/
CM should notify the architect for the inspection of the
precast concrete work. Representatives of the precaster
and the erector participate in this inspection tour and
answer any questions posed by the architect. The GC/
CM should request a final punch list from the architect
so that remedial items can be finished in a timely manner to avoid delaying subsequent trades.
After the precast concrete units have been installed
on the structure in conformance with plans and specifications and the installation is accepted by the architect, subsequent responsibility and liability for the protection of the precast concrete during the construction
phase of the project should rest with the GC. Provisions
for any construction loads that are in excess of stated
design requirements and may occur after precast concrete unit installation are the responsibility of the GC,
not the precaster or erector.

4.1.4.1 Bid Process
Where the selection of a precaster is not negotiated
or controlled by the owner or architect, but is instead
governed by an open-bid situation, the following bid
process is recommended.
STEP 1 — Verification of architect’s concepts and
systems. A review of the proposed precast concrete
concepts during the early design development stage
of the architectural contract documents should be arranged with at least one local precaster. This review
confirms or modifies the architectural concept so that
a realistic design is presented on the bid drawings.
Items to be discussed or reviewed:
• Panelization, form families, piece sizes and weights,
and reveals;
• Shipping and erection issues;
• Architect’s concept for structural support or connections for the precast concrete units so that the architect
can communicate support requirements to the EOR;
• Desired aesthetic issues relative to mixture(s) and
finish(es) and the sample process;
• The architect’s intent for any interfaces with adjacent systems, such as windows, roofing, or building
entrances; and
• Requirements for mockups or other special testing
requirements.

DESIGN

4.1.4 Responsibilities of the General Contractor/Construction Manager / 4.1.5 Responsilbilities of Precaster

STEP 2 — The prebid conference. This is a recommended meeting for all precasters intending to bid the
project, usually held at least three weeks before the bid
date. The design team presents the precast concrete
concepts for the project so that competitive and responsive bids will be obtained. This will improve communications and resolve outstanding questions prior
to preparation of cost estimates and bids. Items to be
discussed include:

• The history of the precaster’s organization as well
as confirmation of the plant’s PCI quality-assurance
(plant certification) program;

• Specifications, PCI plant certification requirements,
and any special provisions;

• Qualifications to the bid that can be listed and
reviewed.

• Design responsibilities and lines of communication;
• The architect’s approved finish samples with information on the mixture proportions, where applicable;
• Prebid submittal requirements, such as proposal
drawings and finish samples;
• Project schedule, shop drawing submittal requirements, and architectural review turnaround times;
• Panelization of precast concrete units;
• Mockups, if applicable;
• Potential problems, discrepancies, or both, found in
the contract documents;
• How and where the project’s precast concrete units
will be structurally attached to the building frame;
• Interfacing with other trades;
• Responsibility for designing, providing, and installing embedded items, anchor bolts, connection
hardware attached to structural steel, bracing, and
other structural items;
• Hardware and reinforcement finishes;
• Special erection needs (access, crane limitations,
and sequence) and logistics; and
• Responsibility for caulking of precast concrete panel
joints.
STEP 3 — Post-Bid scope review: This review allows the architect and GC/CM to review the precaster’s
proposal and confirms the precaster’s ability to satisfactorily meet the project requirements and conform
to design concepts and finish requirements. This material should include:
• Proposal drawings, which express the architectural
precast concrete panelization and structural connection concepts;
• Finish samples;

4

• A list of comparable projects, references, and financial capability;
• Key schedule items, such as mockup panels, shop
drawings and design submittals, mold production,
production start and durations, and erection start
and durations (if applicable); and

If the project allows for a negotiated precast concrete
contract, and the precaster is brought on board during
the initial stages of development, prebid and bid submittal information can be minimized.
Construction coordination. A construction conference should be held at the jobsite after award of the
precast concrete and erection contracts. The GC/CM
should conduct frequent jobsite meetings to coordinate
precast concrete design and erection with the work of
other trades and general building construction.
The coordination meetings should consider all details
of loading, delivery sequences and schedules, types of
transportation, routes of ingress and egress for delivery
trucks and erection cranes, handling techniques and
devices, connections, erection methods and sequences, the effects of temporary bracing on other trades,
and onsite storage and protection. Questions regarding site access, street use, sidewalk permits, oversized
loads, lighting, or unusual working hours should be
addressed at this time.

4.1.5 Responsibilities of the Precaster
Precasters will perform component and connection
design of the members they produce when required
by the contract documents. Precast concrete reinforcement is determined by building codes and industry
standards and the design criteria defined by the contract documents.
All drawings and specifications that convey the requirements for the precast concrete scope should be
provided to the precaster. Pertinent drawings might
include architectural, structural, electrical, plumbing,
and mechanical drawings depending on the size and
scope of the project; approved shop drawings from
other trades; and site plans showing available erection
access and storage areas.

ARCHITECTURAL PRECAST CONCRETE

|

269

4

DESIGN

4.1.5 Responsilbilities of Precaster / 4.1.6 Responsilbilities of Erector

For practical reasons and economy, the precaster first
determines the panelization (panel sizing and joints)
and then the connections. Ideally, a precaster performs
value engineering early in the preliminary design phase
(in a partnering relationship) to reduce construction
costs, improve structural efficiency, facilitate erection
and precast concrete performance.
The precaster should request clarification of ambiguities in writing from the design team through contractual
channels on special conditions not clearly defined by the
design documents. Precast concrete erection and shape
drawings should be submitted to the design team for
approval or acceptance. This submittal is typically done
through the general contractor. When the construction
schedule demands a rapid turn-around time for review
of drawing submittals, the precaster should notify the
design team of their obligations to review and return
submitted drawings within the agreed upon time period
to avoid costly delays in the project schedule. Review
meetings for information exchange and resolution of
conflicts can expedite the approval process.
The precaster prepares detailed shape and erection
drawings and design calculations that are usually signed
and sealed by a professional engineer registered in the
state where the project is located. These drawings and
calculations should show all design criteria, identify all
materials, illustrate precast concrete panel interfacing
with other precast concrete units, the structure and
adjacent materials, and indicate the magnitude and
location of all design loads imparted to the structure
by the precast concrete connections. Design modifications should be permitted only after the design team’s
approval of the proposed change.
The precaster designs the precast concrete panels
and connection hardware for the design loads defined
by the EOR and is responsible for selecting, designing, and locating hardware and panel reinforcement
or items associated with the precaster’s methods of
handling, storing, shipping, and erecting the precast
concrete units. If necessary, this also includes an erection and bracing sequence developed in conjunction
with the erector, EOR, and GC to maintain the stability
of the structure during the erection phase.
Additional design responsibilities for the precaster
should be clearly defined in the contract documents
and may occur when the design team uses Options II
and III (Table 4.1.1). Option III might be used for design-build or with performance specifications.

270

|

ARCHITECTURAL PRECAST CONCRETE

Quality control for product manufacturing is provided
by the precaster according to provisions contained in a
comprehensive quality system manual developed by the
precaster in addition to requirements contained in PCI
MNL-117. Quality assurance is provided through the
precaster’s participation in the PCI Plant Certification
Program. Additional inspection at the owner’s expense
may be required, by specification, through the owner’s
quality assurance agency.

4.1.6 Responsibilities of the Erector
The responsibility for erection of the precast concrete
units may be part of the precaster’s contract, to be performed by the precaster’s own crews or subcontracted
to specialized erection firms, or it may be assigned separately by the GC. Fabrication and erection included in
one contract is recommended by precasters because
this improves coordination and provides single source
responsibility.
Erectors and precasters coordinate development of
efficient connections to facilitate erection for each
project based on their equipment and expertise. The
erector should coordinate the erection plan including
the sequence of erection with the GC/CM and the
precaster.
The precast concrete erector should layout the panels based on the GC/CM’s control lines and elevation
data. This layout should provide panel and joint locations and elevations. This survey should identify any
potential problems caused by building-frame columns,
or beams that are misaligned or out of dimensional
tolerance. Any discrepancies between site conditions
and the erection drawings, which may cause problems
during erection, should be noted in writing and sent
to the GC/CM for resolution prior to the start of erection. Some of these potential problems could include
improper structural steel alignment or hardware installation, errors in bearing elevation or location, and obstructions caused by other trades. Erection should not
proceed until these discrepancies are corrected by the
GC/CM, or until the erection requirements are modified. This survey will also keep the differential variation
in joint widths to a minimum and expedite precast concrete panel erection.
Installation quality assurance will be in accordance
with industry standards, such as the PCI Erectors’
Manual-Standards and Guidelines for the Erection of
Precast Concrete Products (MNL-127). Additional qual-

DESIGN

4.2 Structural Design / 4.2.1.1 Design objectives

ity assurance can be provided by requiring installation
by an industry-qualified or certified erector.

4

4. Loads for plant handling, transportation, and erection, see Section 4.2.9.
5. In-service loads, see Sections 4.2.2 through 4.2.7.

4.2 S TRUCTURAL DESIGN
4.2.1 G
 eneral Considerations
This section constitutes design considerations including a checklist over and above those listed in Chapters
2, 3, and elsewhere in this chapter. The sequence of
these considerations is not of importance, but it is important that the designer use these criteria when making decisions about the ultimate quality and economy
of the final structural system. Design approaches,
determination of loads, dimensioning of precast concrete panels, reinforcement, and contract drawings are
among the subjects treated in detail in this section.
When choosing a façade system it is important to
compare the advantages and disadvantages of both
loadbearing or non-loadbearing units. This is a decision that has unique considerations for each project.
Most architectural panels are non-loadbearing; they
are required to resist only those stresses acting on
the panels themselves. However, the façade is also
capable of acting in conjunction with the rest of the
framing system to carry vertical loads from the floors
and provide stiffness to the structure for lateral restraint in addition to supporting its own weight. This
loadbearing-panel design approach takes advantage
of the high compressive strength of the concrete and
the reinforcement content necessary for handling of
precast concrete units that is available to resist load
in the completed structure. Any increased cost due to
connections and erection time is more than offset by
the reduction of structural framing costs.
For a pronounced horizontal façade, either loadbearing spandrel panels or non-loadbearing façade units
can be used. For façades without a dominant vertical
or horizontal structure, the choice between loadbearing and non-loadbearing façade panels will be governed by the specific conditions of the project.
The typical factors in the structural design of architectural precast concrete units include:
1. Shape and its impact on mold design, see Section 3.3.
2. Properties of the concrete, see Section 3.2.6.
3. Hardware for handling and connections, see
Sections 4.5.4 and 4.5.6.

6. Reinforcement, see Section 4.4.
7. Connections, see Section 4.5
8. Tolerances, see Section 4.6
9. Joints, see Section 4.7

4.2.1.1 Design objectives
Structural integrity of the completed structure is the
primary objective of the structural design. Deflections
must be limited to acceptable levels. The inherent stiffness of architectural precast concrete panels can be
employed to significantly reduce deflections and improve stability of a structure.
Economy is an important design objective when
chosing the structural system since the total cost of a
completed structure is generally the determining factor when comparing alternative construction materials. The designer should attempt to optimize the entire structure and consider the advantages provided by
multifunctional precast concrete panels. In this regard,
the designer should be aware of the major economies offered by standardization or repetition of panel
shapes and sizes. Consideration must also be given to
the cost of large versus small panels regarding weight
limitations on transportation, crane capacity, and crane
location.
Repetition reduces mold construction costs by requiring the construction of fewer molds. Consequently,
production-line processes can be implemented in the
plant enabling a particular casting sequence to be repeated each day, which leads to improvements in efficiency through the repeated operations of familiar
tasks. Handling, storage, and delivery operations are
also simplified with panel repetition, and the risk of errors is reduced. Site efficiency is also improved because
erection sequences can be repeated as well.
Even when a high degree of repetition appears possible, as details are finalized, design discipline may be
required to avoid the creation of a large number of
non-repetitive units. Any budget costs estimated by
the precaster at the initial design stage should take
into account the possibility that the number of different units will increase as the design progresses. If nonrepetitive units must be used, any increase in cost can
be minimized if they can be cast from a master mold

ARCHITECTURAL PRECAST CONCRETE

|

271

4

DESIGN

4.2.1.1 Design objectives / 4.2.1.2 Design criteria

with simple modifications instead of completely different molds. In general, it is relatively simple to alter a
mold if the variations can be contained within the total mold envelope by use of bulkheads or blockouts,
rather than by cutting into the mold surface.
The term “standard” is difficult to define, but panels
cannot truly be described as standard unless they are
identical in every respect. Even relatively minor variations, such as the positions of connections, are sufficient to make a unit non-standard (non-repetitive) and
subject to higher unit costs. However, precasters can
generally accommodate minor panel changes without
incurring significant cost premiums.
The aesthetic design objectives for the structure should
be a matter of concern to the structural engineer. A
precast concrete element or system may achieve all design objectives but the aesthetic ones due to structural
requirements.
For wind or seismic design, the panels should be
designed to limit damage for the extreme events. To
allow the precast concrete panels to undergo less deformation than the supporting structure, the connections can be designed to accomodate the supporting
structure movement. The connections can be designed
to sustain deformations and rotation associated with
extreme design loads without fracture (exhibit strength
and connection ductility).
The direction of the ground motion caused by a seismic event cannot be predicted. Therefore, a structure
shaped to be equally resistant in any direction is the
optimum solution. Experience has shown that a structure that is symmetrical in plan, with minimum torsional eccentricity, generally behaves better in earthquakes
than a structure that is asymmetrical and has its center
of mass and rigidity well separated.

|

The designer of a precast concrete building can
choose to transfer loads through architectural precast
concrete elements or not. In the preliminary design
phase, the structural engineer should recognize their
ability to choose the load transfer mechanism rather
than analyze a predetermined set of criteria.
In some cases, design and analysis of a single precast
concrete element can be completed with very little consideration of other materials and elements in the structure. The weight of the element and the superimposed
loads are simply transferred to the supports, and the
design of the element can be considered independent
of the structure. Occasionally, however, it is necessary
to consider the characteristics of other materials and
elements within the structure. For example, neglecting the relative movement between the precast concrete cladding and the supporting structure may lead
to inaccurate estimates of connection forces. These
movements result from differential volume changes
between the panel and the supporting structure and
deformation of the supporting structure under applied
loads. Forces induced by restrained differential movements between the panel and the supporting structure
are best avoided by design provisions that allow sufficient relative movement at the connections.
Architectural precast concrete design can be considered in three parts:

4.2.1.2 D
 esign criteria

1. Precast concrete elements individually.

Building codes contain design criteria upon which the
design of the building must be based. In some cases,
these criteria are general and it thus becomes necessary for the designer to develop specific criteria. PCI,
the American Concrete Institute (ACI), and other organizations have also developed useful design criteria
and methods.

2. Support system(s) for the precast concrete elements, such as the beam, slab, wall, column, and
foundation.

Selection of design criteria is one of the many important choices that must be made in the design process.

272

In some cases, prescriptive criteria can be used in lieu
of analysis. An example would be the use of a 1/4 in. (6
mm) provision for differential movement between two
adjacent stories in a multistory building as a criterion
for the design of the connections. By adopting this criterion, the designer is making a judgment instead of
calculating the amount of movement for the specific
structure under applied loads. Whenever such general
criteria are adopted, the designer should consider the
limitations involved in their application.

ARCHITECTURAL PRECAST CONCRETE

3. Connections that serve to join the precast concrete
element to its support system.
The design of the architectural precast concrete elements, and the structure of which they are a part, involves load transfer, consideration of stability, and the
potential for movement of the panels or the structure.

DESIGN

4.2.1.2 Design criteria / 4.2.1.3 Checklist

The design engineer is referred to the PCI Design
Handbook—Precast and Prestressed Concrete (MNL120) and Design and Typical Details of Connections for
Precast and Prestressed Concrete (MNL-123) for design
procedures.
The designer must take careful note of the allowable
tolerances for the structural system. This is particularly
important for isolated elements forming a long vertical
line, such as column covers, where any deviation from
vertical is readily noticed.
All non-loadbearing elements should be designed
to accommodate relative movement freely between
the panel and the supporting structural frame and,
whenever possible, without redundant supports, except where provisions are necessary to restrain panel
bowing. It should be noted that high stresses can be
induced if bowing is completely restrained.

4

Detailed Design of Panel – refer to Chapters 3 and 4
1. Concrete mixture proportions.
2. Reinforcement:
a. For function in final position.
b. For stripping, storage, transportation, and
erection.
3. Design of connections, inserts, hardware lifting
hooks, etc. for stripping, storage, transportation,
and erection.
Shop Drawings
For acceptable standards refer to PCI Architectural
Precast Concrete Drafting Handbook (MNL-119),
including:
1. Erection drawings.
2. Anchor layout drawings.
3. Connections details.
4. Piece drawings.

4.2.1.3 C hecklist

5. Hardware details.

The following is a checklist of items the design team
should consider in the design, manufacture, and erection of architectural precast concrete elements.

6. Storage diagrams.

Architectural Requirements
1. Loadbearing or non-loadbearing.
2. Finishes. For full information on the many and
varied types of finish available, contact PCI architectural precast concrete manufacturers or refer to
the PCI Color and Texture Selection Guide.
3. Insulated or non-insulated panels, refer to Section 5.3.
Size of Panel
1. W
 eight limitations.
2. P roduction limitations.
3. T ransportation weight and dimension limitations.
4. E rection feasibility and access.
5. S tress limitations.
Supporting Structure
1. P oints of load application.
2. O
 verall stability.
3. S tability during erection.
Standardization
Obtain maximum repetition of similar units to reduce
overall costs.
Design of Connections
Refer to Section 4.5.

7. Drawings for special handling.
Molds
For information on the various materials used for the
manufacture and design of molds, contact a PCI certified architectural precaster.
Production
For information regarding materials to be used and
production methods to be followed, refer to PCI
Manual for Quality Control for Plants and Production
of Architectural Precast Concrete Products (MNL-117).
Quality Control
For information on the requirements of quality control over the entire production sequence, refer to PCI
MNL-117.
Transportation
Refer to Section 2.1.4 in PCI Erectors’ Manual–
Standards and Guidelines for the Erection of Precast
Concrete Products (MNL-127) for:
1. Types of trailers.
2. Types of support frames.
3. Support material.
For further detailed information on methods, materials, and equipment used in handling and transporting
all types of precast concrete units, contact a PCI certified architectural precaster

ARCHITECTURAL PRECAST CONCRETE

|

273

4

DESIGN

4.2.1.3 Checklist / 4.2.2 Determination of Loads

Erection
For the successful use of architectural precast concrete,
the designer must envision the erection process, refer
to PCI Erectors’ Manual-Standards and Guidelines for
the Erection of Precast Concrete Products (MNL-127):

ing handling, transportation, and erection (prior
to installation on the building). (Design for these
loads is normally the responsibility of the precaster,
as the stresses during handling and erection often
govern the design of the precast concrete units.)

1. U
 nimpeded access to allow for continuous erection, working and storage space, staging area for
trailers and cranes.

3. Those in service structural loads that are externally
applied to the unit or transferred to the unit by the
behavior of the supporting structure. These loads
can include wind, seismic, snow, blast, floor live
and dead loads, or construction loads, and should
be shown on the contract drawings. In-service
loads are set by the governing building code and
are multiplied by the appropriate load factors for
use in design.

2. E quipment, cranes, monorail hoists, etc.
3. L ifting, turning, tilting units.
4. S urvey of new or existing structural frame, location
of cast-in hardware.
Economy
1. D
 esign, production, transportation, and erection
costs.
2. P rogress payments for completed units stored at
plant.
3. E conomy of precast concrete.
Tolerances – refer to Section 4.6
1. D
 esign of project clearances.
2. C
 ast-in-place concrete or steel support structure
tolerances, sway, creep, differential deflection.
3. P roduction tolerances, dimensions, bowing, warping.
4. E rection tolerances:
a. B
 etween precast concrete units and supporting
structure.
b. B
 etween precast concrete units.

4.2.2 D
 etermination of Loads
The structural design of the architectural precast concrete elements involves load transfer and consideration
of stability. Achievement of structural design objectives
also requires the consideration of movement or the potential for movement within the system of which the
precast concrete element is a part.
The forces that must be considered in the design of
structures that contain architectural precast concrete
can be classified as follows:
1. Those caused by the precast concrete members
(for example, their self weight and effect on seismic forces). Each unit has to support its own weight
and may have to transmit this and the weight of
other units or elements, such as windows, to the
structure.
2. Loads imposed on the precast concrete units dur-

274

|

ARCHITECTURAL PRECAST CONCRETE

4. Loads resulting from restrained volume change or
support-system movement. These forces are generally concentrated at the connections and sometimes govern the connection design. The designer
should provide simple load paths through the connections, and ductility within the connections. This
will reduce the sensitivity of the connection to these
forces and the necessity to precisely calculate loads
and forces from, for example, volume changes and
frame distortions. The number of load transfer
points should be kept to a practical minimum. It is
recommended that no more than two connections
per panel be used to transfer gravity loads.
The forces or stresses imposed on precast concrete
units during the manufacturing-erection processes occur primarily as a result of the member being in an
orientation differing from that of its final position in
the structure. Manufacturing and erection stresses are
controlled by the fact, for example, that the concrete
strength at time of stripping may only be a fraction of
its final design strength. Thus, for maximum economy
of material, the production process should be given
consideration as part of the structural design process. In
particular, architectural precast concrete panels should
be shaped such that they will be sufficiently stiff in the
direction of handling-induced stresses. Limitations on
product dimensions imposed by transportation should
also be considered during the design process; the designer should be familiar with legal load limitations
and cost premiums associated with transporting overheight, over-width or over-length members.
In-service loads may not be as critical as those imposed during the manufacturing-erection process, except for panels in zones of high seismicity, hurricane

DESIGN

4.2.2 Determination of Loads / 4.2.3.2 Shrinkage

winds, or loadbearing panels. Generally, the determination of in-service loads, including loads imposed on
the structure by the architectural precast concrete are
the responsibility of the EOR.
In high-rise construction, vertical precast concrete
panels can span multiple floors. Multiple vertical floor
spans require gravity loads to be supported at only
one floor per panel thus loading every second or third
floor-see Fig. 2.4.14, page 70. That way, most floors
can be designed without the need to support the gravity loads of the exterior skin, thus reducing the overall
structure’s cost. This approach may not be practical in
seismic zones because of drift requirements.

4.2.3 V
 olume Changes
Volume changes of precast concrete are caused by
variations in temperature, shrinkage due to air-drying,
and creep caused by sustained stress. If precast concrete members are free to move or deform, volume
changes cause little or no stress in the panel. However,
if the panel is restrained, significant stresses and cracking may develop. For example, welding long precast
concrete panels directly to the structure at both ends
should be avoided. Different temperatures on the interior and exterior of the building may cause the panel
to bow. This bowing can be resisted by center connections, but this causes stresses in the panel that should
be considered in design.
Volume changes due to temperature variations can
be positive (expansion) or negative (contraction), while
volume changes from shrinkage and creep are only
negative. The amount of movement anticipated due to
volume change must be determined to properly design
joints and connections.
Because architectural precast concrete members are
generally not subjected to large sustained stresses, volume changes due to creep are usually minimal. Thus,
the volume changes in architectural precast concrete
that must be recognized and accounted for are due
to temperature and drying shrinkage. The amount
of these volume changes that are tolerable depends
on jointing and connection details of the structure. In
most cases, any panel shortening that takes place prior
to making the final connections will reduce the shrinkage and creep strains to manageable proportions.
For low- to medium-rise structures the major effect
of volume change in the precast concrete units will
be in the horizontal direction. Nevertheless, vertical

4

elements, such as loadbearing wall panels, are also
subject to volume-change strains. Volume change effects in the vertical direction will usually be significant
only in high-rise buildings, and then only differential
movement between elements will significantly affect
the performance of a structure. This can occur, for example, at the corner of a building where loadbearing
and non-loadbearing panels may meet or where precast concrete panels are connected to a cast-in-place
frame.
Estimates of building movement must be tempered
with engineering judgment. Floors and interior walls
attached to exterior loadbearing panels will tend to
restrain vertical movement. Also, heavily loaded elements will tend to distribute load to less heavily loaded
ones.

4.2.3.1 Temperature effects
The coefficient of thermal expansion of concrete varies with the aggregate type. Ranges for normalweight
concrete are 5 to 7 x 10-6 in./in./°F (9 to 12.6 x 10-6
mm/mm/°C) when made with siliceous aggregates and
3.5 to 5 x 10-6 in./in./°F (6.3 to 9 x 10-6 mm/mm/°C)
when made with calcareous aggregates. The approximate values for structural lightweight concretes are
3.6 to 6 x 10-6 in./in./°F (6.5 to 10.8 x 10-6 mm/mm/°C),
depending on the type of aggregate and amount of
natural sand. Coefficients of 6 x 10-6 in./in./°F (10.8 x
10-6 mm/mm/°C) for normalweight and 5 x 10-6 in./in./
°F (9 x 10-6 mm/mm/°C) for sand-lightweight concretes
respectively, are frequently used. If greater accuracy is
needed, tests should be made on the specific concrete
to be used in the project.
Because the thermal coefficient for steel is approximately 6 x 10-6 in./in./°F (10.8 × 10-6 mm/mm/°C) , the
addition of steel reinforcement does not significantly
affect the concrete coefficient.

4.2.3.2 Shrinkage
Precast concrete members are subject to air-drying as
soon as they are removed from the molds. During this
exposure to the atmosphere, the concrete slowly loses
some of its original water causing shrinkage to occur.
About 40% of drying shrinkage occurs by age 30 days
and about 60% by age 90 days. The rate and amount
of shrinkage is dependent on the concrete mixture
proportions and materials, the temperature and hu-

ARCHITECTURAL PRECAST CONCRETE

|

275

4

DESIGN

4.2.3.2 Shrinkage / 4.2.4 Design Considerations for Non-Loadbearing Wall Panels

midity of the environment, and the size and shape of
the member.
During the first year, the total unit length change due
to drying shrinkage of normalweight concrete typically
ranges from about 4.0 to 6.5 x 10-6 in./in. when exposed to air at 50% relative humidity (RH). Lightweight
concrete containing all natural sand fines has one-year
shrinkage values that range from 5.5 to 9.0 x 10-6 in./
in.; concrete with a unit shrinkage of 6.0 x 10-6 in./in.
shortens about 0.72 in. (18 mm) per 100 ft (30 m)
while drying from a moist condition to a state of moisture equilibrium in air at 50% RH. In comparison, this
equals approximately the thermal contraction caused
by a decrease in temperature of 100 °F (38 °C).
Differential shrinkage within panels containing face
mixtures and/or material prone to shrinkage (brick,
tile, etc.) must be carefully evaluated to avoid excessive
bowing.
Sufficient reinforcement must be used in each unit
to control the distribution and crack widths of any
shrinkage cracking. Where units have complex shapes,
and particularly where they have unbalanced volumes,
unsymmetrical reinforcement, large protrusions, or
changes of section, the risk of shrinkage cracking is
increased. Distortion (bowing or warping) of the panel
can also occur due to these causes.

4.2.3.3 C reep
When concrete is subjected to a sustained load, the
deformation may be divided into two parts: (1) an
elastic deformation that occurs immediately, and (2) a
time-dependent deformation (creep) that begins immediately upon application of load or prestress and
continues over time.
For design, it is convenient to refer to specific creep,
which is defined as the creep strain per unit of sustained stress. The specific creep of architectural precast
concrete panels made with normalweight aggregates
per unit stress (psi) can range from 0.5 to 1.0 x 10-6
in./in./psi. About 40% of the creep occurs within 30
days of load application, and about 60% occurs within
90 days.

4.2.4 D
 esign Considerations for
Non-Loadbearing Wall Panels
Non-loadbearing (cladding) panels are those precast
concrete units that transfer their own dead loads and

276

|

ARCHITECTURAL PRECAST CONCRETE

any imposed dead loads, such as windows, to the structural frame or foundation. They are designed to resist
wind and seismic forces and also may be designed to
take wind or seismic loads distributed to the panel
from elements fastened directly to it, such as windows.
The forces imposed during manufacturing and erection also need to be considered in the design of the
panels. Wind may control the design loads in hurricane
regions, building corner zones, or panels adjacent to
large openings. The forces resulting from seismic considerations will generally govern connection design,
but will usually result in panel stresses less than those
imposed during manufacturing and erection.
All non-loadbearing panels should be designed to
accommodate some movement. Potential differential
movements between the panel and the supporting
structure must be evaluated, and care taken to prevent unintended restraints from imposing additional
loads on the panel. The supporting structure may deform due to the weight of the panel, volume changes
in concrete frames, or rotation of supporting beams.
Unintended load transfer among adjacent panels
should be avoided. This is accomplished by detailing
joints with sufficient space so that the anticipated deformation of the supporting structure and the panel
will be less than the space between elements. Also, the
connections must be designed and installed to permit
expected deformations to freely occur.
The panel cross-section is generally chosen for architectural or aesthetic reasons. A panel that is too thin
may bow or deflect excessively, thereby creating caulking problems at the building corners or fit-up and leakage problems at attached windows.
As the height and length of a building increase, the
cumulative movements at the top or ends of the structure increase. The movements of exterior walls can affect the interior partitions on upper floors resulting in
distress to, or cracking of, the partitions. Non-structural
components at the building interior must be detailed
to allow for volume change movements of exterior
precast concrete structural walls. If non-structural elements, such as drywall ceilings, or interior drywall or
masonry partitions are attached to wall panels unrestrained against bowing, those items should be attached with “soft” or flexible connections.
Where loadbearing and non-loadbearing panels meet
(for example, at the corners of a building) it is possible

DESIGN

4.2.4 Design Considerations for Non-Loadbearing Wall Panels / 4.2.4.1 Deformations

that differential volume change movements may occur.
If connections are employed that restrain these movements, it is likely that the connections will attempt to
transfer significant vertical forces. Structural behavior
of the building at corners where panels meet requires
specific attention and should be designed for volume
change movement forces as well as for other structural
forces.
Design consideration of panels meeting at corners
should include the influence of temperature differentials due to sun exposure on the panel connections.
Depending on the exterior panel color and plan orientation of the building, 9 to 14 °F (5 to 8 °C) temperature differentials may develop in elements that are in
direct sun versus those that are shaded.

4.2.4.1 D
 eformations
Deflection of a panel support is a function of the
stiffness of the support. Where adjacent panels are
supported on different portions of the building frame
with differing stiffnesses, relative deflections between
adjacent panels may occur. This is often the case at
building corners, where the structural arrangement
may result in significantly different support stiffnesses.
It is also a concern where the structural frame cantilevers. If panels are attached in a manner that tends to
prevent relative displacement, developed panel stresses
must be evaluated.
The weight of a series of small panels supported on a
long span flexible beam is shown in Fig. 4.2.1. The support beam will deflect (or rotate) in increments as each
panel is erected, resulting in an in-plane rotation of the
panels previously erected, which may cause unintended

4

restraint forces to develop in the longitudinal direction.
Alternatively, this problem could be solved by providing a single bearing connection for each panel or connecting panels at or close to columns. The EOR should
design the supporting beam with minimal vertical deformation so that the precast concrete erector can erect
the panels to be level and at the correct final elevations
without having to reset or realign panels (see also discussion in Section 2.5 on wall-supporting panels).
Loading from other sources may also cause a deflection-related problem. For example, if precast concrete
is erected prior to floor slab construction, the weight
of the floor may deflect the support beam and cause
a problem similar to the one shown in Fig. 4.2.1. The
connections should be designed to allow the supporting beam to deflect, but the beam should be stiff
enough that panel joint widths remain within the specified tolerances.
Because of the difficulty in detailing these connections
and erecting panels on a flexible support, it is generally
preferable to provide panels that span from column to
column. In the case where the wall panels or spandrels
are supported by the columns, the primary concern for
deflection is the control of cracking and potential distress in the joint between the wall finish and the floor
finish. Because the wall and floor (or roof) are supported independently, they are loaded and deflect independently. The acceptable range of movement depends on
the finishes, but, for example, typical office occupancy
finishes can tolerate vertical differential movement on
the order of 1/4 to 1/2 in. (6 to 13 mm). The appropriate
load to use in calculating this deflection is 50% of the
design live load. Lastly, the cladding tieback connec-

Fig. 4.2.1 Deformation of panels on flexible beam.
Separate panels

Single panel

Deflected position of supporting member due to weight of panels

ARCHITECTURAL PRECAST CONCRETE

|

277

4

DESIGN

4.2.4.1 Deformations

tions must be detailed to accommodate the range of
vertical movement at the building perimeter.
In the case of relatively light cladding, that is, when
the load from the panel plus window system is less
than or equal to 25% of the total load on a steel spandrel support beam, the following limits apply. The recommended limit on steel spandrel beam dead load deflection prior to erecting the cladding are span length
(in inches) divided by 480 with an absolute limit of 3/8
in. (10 mm). The loads appropriate to this calculation
are those in place prior to cladding. The deflection due
to all dead loads including the cladding plus window
system should be limited to span length divided by 480
with an absolute limit of 5/8 in. (16 mm).
The deflection limits for live load are based on the
movement allowed by the detailing of the completed
cladding plus window system.
Although large movements could theoretically be accommodated, common detailing practice would allow
for movements in the range of 1/4 to 1/2 in. (6 to 13
mm). These deflections would be the net allowable
movement after accounting for the dead load placed
after the cladding and window system is completed. In
relative terms the live load deflection should be limited
to span length divided by 360. The portion of design
live load used depends on the expectation of its presence and could vary between 50 and 100% of the total design live load.
When the panel and window system weight exceeds
25% of the total dead load on the steel spandrel beam,
different limits apply. In this case, the critical dead load
deflection is due to the dead load in place at the time of
cladding plus the dead weight of the wall system itself.
The deflection due to these loads should be limited to
span length divided by 600 with an absolute limit of 3/8
in. (10 mm).
The deflection limits for dead loads imposed after
cladding and live loads are the same as for the case of
relatively light cladding. They are:
•A
 dditional dead load: span length divided by 480
with an absolute limit of 5/8 in. (16 mm).
• Live load: span length divided by 360 with an absolute limit in the range of 1/4 to 1/2 in. (6 to 13 mm).
It should be noted that these limits are set by the
allowable movements in the joints and connections.
Thus, for the most part, camber of the steel beam

278

|

ARCHITECTURAL PRECAST CONCRETE

is not a means to address these concerns. The only
exception would be deflection prior to installation of
cladding. However, these loads and deflections are
relatively small in magnitude and probably would not
justify cambering the beams.
Bowing due to temperature differences between the
inside and outside of a wall panel is the most prevalent cause of panel deformation after placement of the
precast concrete panel in the structure. Panels generally will bow outward. If supported in a manner that
will permit bowing, the panel will not be subjected to
stress as a result of the bowing. However, if the panel
is restrained laterally at the middle, bowing restraint
stresses in the panel and forces in the structure will
occur.
Non-loadbearing panels that contain openings, such
as window panels, may develop stress concentrations
at these openings, resulting from unintended loading
or restrained bowing. Figure 4.2.2 illustrates the profile of a panel that tends to deflect outward due to
warming of the exterior surface. Experience indicates
that if the panel is restrained on all four edges, hairline cracking radiating from the corners of the opening
may develop. While these stress concentrations may be
partially resisted by reinforcement, the designer should
always consider methods of eliminating imposed restraints. Good design practice requires that areas with
abrupt changes in cross-section be reinforced and
should be rounded or chamfered whenever possible.
Moisture differences between the inside and outside
of an enclosed building can also cause bowing; however, the calculation to determine the magnitude of bowing from this effect is much less precise and involves
more variables. The exterior layer of the concrete panel
absorbs moisture from the atmosphere and periodic

Fig. 4.2.2 Cracking due to restrained bowing.

DESIGN

4.2.4.1 Deformations

precipitation, while the interior layer is relatively dry,
especially when the building is heated. This causes the
inside layer to shrink more than the outside, causing an
outward bow. The outward shrinkage bowing would
tend to balance the theoretical inward thermal bowing
in cold weather, which is believed to explain the observation that “wall panels always bow out.”
While the magnitude of bowing is usually not significant, in the case of wall panels it may cause unacceptable separation at the corners (Fig. 4.2.3) and
damage to joint sealants. It may therefore be desirable
to restrain bowing in these locations with one or more
connectors between panels.
Fig. 4.2.3 Corner separation due to thermal bow.

Non-loadbearing panels should be designed and installed so they do not restrain frames from lateral translation. If such restraint occurs, the panels may tend
to act as unintended shearwalls (significant diagonal
compression may occur) and become overstressed.
To prevent this, panels that are installed on a frame
should be connected in a manner to allow frame distortion (connected at the top and bottom only and left
free along the sides). In some cases, especially in high
seismic regions, special connections that allow movement may be required. The space between the panel
and the supporting frame required for erection will
usually be sufficient to prevent contact during lateral
deformation of the frame.
Frame shortening of buildings and its affect on precast concrete panels is only of concern if it produces a
differential movement between the building cladding
and the structure. There is little cause for concern if
the building and the precast concrete cladding move
approximately the same amount. However, this seldom
occurs.

4

For low-rise and mid-rise buildings, column shortening is of little concern. For high-rise buildings, the column shortening related to the structure’s dead load often takes place before the cladding system is installed.
Any shortening due to cladding loads and floor live
loads should be determined and documented, and
evaluated in the context of the panel connections and
joint details. Consideration should also be given to the
timing of the erection process. If column shortening is
a design consideration, it should be determined and
documented by the building’s structural engineer.
The vertical shortening of concrete columns should be
considered when structures are tall. A 40-story, cast-inplace concrete building can shorten as much as 5 to
6 in. (125 to 150 mm) during and after construction.
Each of the floors exhibits a proportional shortening.
These shortenings are cumulative for the height of the
structure. At each level, the differential shortening between two adjacent floors is a small amount that can
be accommodated by the cladding panels. At the lowest level, if the panel is rigidly supported at the base
(such as a foundation or transfer girder) and the panels
are stacked to support the wall above and tied into
the structure, the gradual shortening of the structure
above may induce unintended loading on the panel.
In such cases, the panel connections and joints should
be designed to permit the calculated deformation. The
full shortening of concrete columns may take several
years, although a major proportion occurs within the
first few months after construction. Frame shortening
must be considered to determine true fabrication elevations. Unless this is done, panel heights may not
match the structural frame and connections will not
line up. Panel connections must line up at the time
of erection and there must be sufficient space at the
joints to compensate for future movement.
A similar design situation will occur when two adjacent columns have significantly different loads. For example, the corner column of a structure will usually be
subjected to a smaller load than the adjacent columns. If
both columns are the same size (as is often the situation
for architectural reasons) and reinforced approximately
the same, they will undergo different shortening.
If adequate clearance is not provided between the
precast concrete panels and the support structure, or
if the connections do not allow for unrestrained movement, loads from adjacent floors can be imposed on
non-loadbearing panels. These loads can cause ex-

ARCHITECTURAL PRECAST CONCRETE

|

279

4

DESIGN

4.2.4.1 Deformations / 4.2.4.2 Column covers and mullions

cessive stresses at the “beam” portion of an opening
(Fig. 4.2.4). This condition can be prevented by locating connections away from critical sections. Unless a
method of preventing load transfer, such as slotted
angle connections, can be developed and permanently
maintained, the “beam” should be designed for some
of the floor loads. Determining the magnitude of such
loads requires engineering judgment.

4.2.4.2 C olumn covers and mullions
The use of precast concrete panels as covers over steel
or concrete columns and beams is a common method
of achieving architectural expression, special shapes, or
fire rating and specific finishes in an economical manner. Column covers and mullions are usually supported
by the structural column or the floor, and are themselves designed to transfer no vertical load other than
their own weight. The vertical load of each length of
column cover or mullion section is usually supported
at one elevation, and tied back at the top and bottom
to the floors for lateral load transfer and stability. In
order to minimize erection costs and horizontal joints,
it is desirable to make the covers or mullions as long
as practical, subject to limitations imposed by weight,
handling, and story drift.
Connections must allow for relative horizontal movement between floors. This may cause the column covers
to rock back and forth between bearing connections. The
length of a column cover will be dependent on transportation and lifting limitations, architectural considerations,
and the ability of the structural column to support a specific concentrated load (panel weight) locally.
Mullions are vertical elements serving to separate
glass areas. Because mullions generally resist wind

loads applied from the adjacent glass, they must be
stiff enough to maintain deflections within the limitations imposed by the window manufacturer. With
these thin flexible members, consideration should be
given to prestressing to prevent cracking.
Column covers and mullions are usually major focal
points in a structure, and aesthetic success requires
that careful thought be given to all facets of design
and erection. The following are some items that should
be considered:
1. Because column covers and mullions are often
isolated elements forming a long vertical line, any
variation from a vertical plane is readily observable.
This variation can be the result of the tolerances
allowed in the structural frame. To some degree,
these variations can be handled by precast concrete
connections with adjustability. The architect should
plan adequate clearance between the panel and
structure to allow the tolerances of the structural
frame to be accommodated. For steel columns, the
architect should consider the clearances around
splice plates and projecting bolts.
2. Gravity support should be provided by two bearing points at only one elevation and connections
at additional locations for lateral loads and stability. When access is available, consider providing an
intermediate connection for lateral support and
restraint of bowing.
3. Column covers and mullions that project from the
façade will be subjected to shearing wind loads. The
connection design must account for these forces.
4. Members that are exposed to the environment will
be subjected to temperature and humidity change.
Horizontal joints between abutting precast con-

Fig. 4.2.4 Unanticipated loading on a non-loadbearing panel.

Floor load imposed due
to unintended restraint
unless slotted connection
is used to allow vertical
movement

Section

280

|

ARCHITECTURAL PRECAST CONCRETE

Flexure of beam

Elevation

DESIGN

4.2.4.2 Column covers and mullions / 4.2.5 Design Consideration for Loadbearing Wall Panels

crete column covers and mullions should be wide
enough to permit length changes and rotation
from temperature gradients. The behavior of thin
flexible members will be improved by prestressing.

The design and structural behavior of exterior architectural precast concrete bearing wall panels is dependent upon the panel shape and configuration, and
should consider the following:

5. Due to vertical loads and the effects of creep and
shrinkage, cast-in-place concrete columns will
tend to shorten. The width of the horizontal joint
between abutting column covers and mullions
should be sufficient to permit this shortening to
occur freely as well as handle rotation from temperature gradients.

1. Gravity loads and the transfer of these loads to the
foundation. Vertical (gravity) loads are parallel to
the plane of the wall at an eccentricity influenced
by the geometry of the wall, location of load, and
manufacturing and erection tolerances.

6. The designer must envision the erection process.
Column cover and mullion connections are often
difficult to reach and, once made, difficult to adjust.
The difficulty of access is compounded when all four
sides of a column are covered by several column
cover units to obtain the full height. Sometimes this
condition can be solved by welding the lower piece
to the column and anchoring the upper piece to the
lower with dowels, or by a connection that does not
require access.
7. Insulation may be placed on the interior face of
the column cover, or alternatively, it may be applied to the structural column directly. Such insulation will reduce heat loss at these locations and
also minimize temperature differentials between
exterior columns and the interior of the structure.
Connections details must be chosen to accommodate either choice.
8. Column covers or mullions can be combined with
adjacent spandrels to minimize number of units
and joints. Alignment of rustications on the column
covers with the horizontal details on the spandrel
panel is facilitated.
9. Where uniformity of architectural finish is required
on two or three sides of column covers, the designer should be guided by a precaster regarding
the feasibility of this requirement. For example, to
ensure uniformity of finish, it may be necessary to
segmentally cast the units. Consultation with a
precaster is recommended.

4.2.5 D
 esign Considerations for
Loadbearing Wall Panels
Most of the items for non-loadbearing wall panels
also must be considered in the analysis of loadbearing
wall panels.

4

2. Magnitude and distribution of lateral loads perpendicular to the plane of the wall (wind and seismic) and the means for resisting these loads using shearwalls and floor diaphragms. Loads in the
horizontal direction may be both parallel to and
perpendicular to the plane of the wall. For typical
precast concrete structures, improved redundancy
and ductility are achieved by connecting members
into a load path to the lateral-load-resisting system. The load path in the lateral-load-resisting system must be continuous to the foundation.
3. Location of joints to control volume change deformations due to concrete creep, shrinkage, and
temperature movements; influence upon design
for gravity and lateral loads; and effect on nonstructural components. The volume change effects
will usually only be significant in high-rise buildings,
and then only differential movements between elements will significantly affect performance of the
structure. This can occur, for example, at the corner of a building where loadbearing and non-loadbearing panels meet or at re-entrant corners.
4. Connection concepts and types of connections required to resist the various applied loads. In some
cases, local practice may suggest one type of connection over another, for example, use of bolts
rather than welds. Welded connections need to be
accessible to allow efficient welding.
5. Tolerances required for the structure being designed
with regard to production and erection for both
precast concrete units and connections, including
tolerances for interfacing different materials.
6. Specific design requirements during the construction stage that may control designs, such as site
accessibility.
The design of exterior walls using loadbearing architectural panels does not differ from two-dimensional
frame design, once the panel is isolated and taken as
a free body. Accepted design procedures and code re-

ARCHITECTURAL PRECAST CONCRETE

|

281

4

DESIGN

4.2.5 Design Consideration for Loadbearing Wall Panels / 4.2.6 Design Considerations for Non-Loadbearing Spandrels

quirements apply to the design of the wall. Perhaps
the only design consideration difference is recognizing
the role precast concrete panel production and erection play in the overall design process. Similarly, usual
accepted procedures and code requirements apply to
the design of an individual precast concrete panel and
its components.

Fig. 4.2.5 Horizontal and vertical rib panels.

Wall panel size and shape can be affected by the details and locations of the vertical and horizontal panelto-panel connections. Both gravity load transfer between panels and gravity and axial load combinations
caused by lateral loadings or size of window openings
can become the major factors influencing panel structural dimensions and connection design. Although,
for most precast concrete exterior bearing-wall structures, it will be found that the gravity dead and live
load condition will control structural dimensions of a
panel rather than load combinations that include lateral loads. Portions of walls with large openings may
be subjected to significant axial loads. These portions
may require reinforcement with closely spaced ties.
Often the mullion size of the panel will not be controlled by the minimum concrete area required by design. Rather, minimum dimensions for grouting panels at horizontal joints and for placing reinforcement,
space for locating handling devices, or space required
to accommodate a variety of connection conditions
may determine mullion sizing.
Panels may be designed to span horizontally between
columns or vertically. Whether or not the architectural
panel of the exterior wall is placed horizontal or vertical
depends primarily on the methods or details selected for
transferring loads and making connections. When spanning horizontally, they are designed as beams or, if they
have frequent, regularly spaced window openings as
shown in Fig. 4.2.5(a), as Vierendeel trusses. If a large portion of the panel is window opening, as in Fig. 4.2.5(c),
it may be necessary to analyze it as a rigid frame. When
the panels are placed vertically, they are usually designed
to be similar to columns. Because of the large height-tothickness ratios and the magnitude and eccentricity of the
loads, the in-place stresses may control the design.
A horizontal Vierendeel truss–type panel lends itself
to simple handling because it can be shipped in its
erected position. It requires vertical load transfer connections at each story level, and requires only minimal
erection handling and bracing.

282

|

ARCHITECTURAL PRECAST CONCRETE

(a) Truss-type panel

(b) Channel-type panel

(c) Window-mullion panel

A two-story vertical panel requires additional design
considerations for erection handling because it needs
to be rotated from its shipped postion during erection.
Figure 4.2.5 shows architectural wall panels, generally
used with relatively short vertical spans (although they
may sometimes span continuously over two or more
floors).
When stemmed floor or roof members are used, the
width of loadbearing walls or spandrels should module with the double-tee width. In other words, for 12
ft (3.7 m) double tees, walls should be 12, 24, or 36
ft (3.7, 7.3, and 11 m) wide. Local precast concrete
producers should be contacted for the availability of a
particular module.
Dimensions of architectural panels are usually selected based on a desired appearance. When these panels
are also used to carry loads, or act as shearwalls, it is
important to have some engineering input in the preliminary architectural design stages of the project.

4.2.6 Design Considerations for
Non-Loadbearing Spandrels
Non-loadbearing spandrels are precast concrete ele-

DESIGN

4.2.6 Design Considerations for Non-Loadbearing Spandrels

ments that are less than story height, made up either as
a series of individual units or as one unit extending between columns. Support for the spandrel weight may
be provided by either the floor or the columns, and
stability against eccentric loading is usually achieved by
connections to the floor or the column (Fig. 4.2.6).
Spandrels are usually part of a window wall; therefore, consideration should be given to limiting the
vertical deflections and rotations of the spandrel to
that consistent with the requirements of the window
manufacturer. Spandrels are also commonly used as
vehicle-impact restraints in parking structures in addition to providing perimeter design features. Doing
so eliminates the need for an upturned cast-in-place
beam or cable system.
For spandrels that extend in one piece between columns (usually 20 to 60 ft [6 to 18 m]), it is preferable
that the gravity supports be located on or near the columns. This arrangement will minimize interaction and
load transfer between the floor and spandrel, which
reduces the structural framing costs. Column sections
are generally very stiff vertically, resulting in minimal
deformations, and can be readily designed to resist the
eccentricity of panel weight.
While structural steel frames tend to have good dimensional control, rotation and deflection as controlling parameters of the design for the steel support beams are

4

much more common than for concrete framing. The
structure should be designed to resist torsional rotation,
due to eccentric loadings on perimeter beams by the EOR.
The steel frame designer must consider this flexibility and
design for the forces transferred to the frame through
the connections. The EOR should clearly specify connection points when flexibility is of concern. Otherwise, the
precast concrete design engineer will need to verify with
the EOR rather than assume that the structure has sufficient rigidity to allow spandrels to be erected without
subsequent realignment and to allow deformations to be
limited to those specified for the spandrels.
When precast concrete spandrel panels are supported on edge beams, it is desirable to design gravity connections to transmit the loads to the beam centerline.
This is often not practical because of the large resulting
eccentricities created and the large connections that
are difficult to conceal. It is often preferable to provide
gravity supports for the precast concrete panels at the
columns. It is preferable to cantilever a slab or steel
bracket from the side of the panel support beam to
allow gravity support close to the back of the spandrel.
This may cause torsion in the support beam that must
be recognized from both a strength and stiffness perspective. Similarly, when tieback connections are made
to the underside of a steel beam, the frame designer
(EOR) should make provision for torsional loading by
specifying heavier members, braces, or gussets. Tie-

Fig. 4.2.6 Forces on a spandrel panel.
Wind or seismic pressure/
suction from window

Window
Tieback
force

Spandrel weight

Spandrel
weight

Tieback force
Wind or
seismic pressure/
suction on spandrel

Spandrel weight

Tieback force
Spandrel support
Tieback force
Spandrel support

Spandrel
support

Tieback force

Tieback
force
Wind or seismic pressure/
suction from window

Floor Connection

Column Connection

Roof Connection

ARCHITECTURAL PRECAST CONCRETE

|

283

4

DESIGN

4.2.6 Design Considerations for Non-Loadbearing Spandrels

backs to columns may require stiffeners be added to
resist local bending. The EOR needs to understand the
effects of localized tieback connections so those stiffeners can be detailed and supplied with the structural
steel.
If the deflection of the structural frame is sensitive
to the location or eccentricity of the connection, limits
on connection eccentricity should be given on contract
drawings and also shown on the erection drawings.
This is particularly important for heavy members bearing on light members, such as open web joist or cantilevered structural members.
The weight of a series of wall panels supported on
a flexible beam will cause deflection (or rotation) of
the edge beam. To prevent imposing unintended loads
on the panel, the connections must be designed and
installed to permit these deformations to occur freely.
The advantages of wall panels spanning between columns (see Section 2.5) become apparent in steel structures, where these units may be supported directly off
the columns. They may also help to balance the load
on the columns.
Consideration should also be given to spandrels that
are supported at the ends of long cantilevers. The EOR
must determine the effects of deflection and rotation
of the support, including the effects of creep, and arrange the details of all attachments to accommodate
this condition (Fig. 4.2.7). A particularly critical condition
can occur at the corners of a building.
When panels supported on cantilevers are adjacent
to panels supported in a different manner, differences
in deflection of the supporting structure may result in
joint tapers and jogs in alignment. The possibility of increased deflection and rotation of the panel over time,
resulting from creep of the supporting cantilever, must
also be considered.
Connections may be detailed to allow for final adjustment after initial erection. However, normal erection procedures assume a panel can be set and aligned
without returning for later adjustment. Often the best
way to deal with this condition is to use a support
scheme that does not rely on cantilever action (Fig.
4.2.8[a & b]). The bracing supports in Fig. 4.2.8 are
normally supplied and erected by the structural steel
or miscellaneous steel subcontractor prior to precast
concrete erection.

284

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 4.2.7 Effect of cantilever supports.

Window
Spandrel

Possible support
rotation
Cantilever
deflection
(Elastic plus creep)
Beam or slab
Window

Fig. 4.2.8(a) & (b) Bracing support.

Support member

Bracing
support

(a)

Structural support
members

(b)

DESIGN

4.2.6 Design Considerations for Non-Loadbearing Spandrels / 4.2.7 Design Considerations for Loadbearing Spandrels

Spacing of lateral supports for a spandrel should not
exceed 50 times the least width of compression/flange
or face. However, the spans of non-loadbearing spandrels on parking structures have frequently exceeded
50 times the width of the top of the member, and no
problems have been observed. This is undoubtedly because they typically carry only their own weight, which
is concentric. Where lateral (vehicle impact) loads are
applied to the spandrel, lateral support of the spandrel
into the deck are typical.
Because the contract bid documents (drawings and
specifications) are the first line of communication between the designers of record and the precast concrete
design engineer, it is imperative that the documents
clearly indicate the connection concept. For example, if
spandrel loads are to be resisted by the columns rather
than the floor beams, it must be indicated on the contract documents. This intent can be shown schematically and further described in applicable notes. Also,
the division of responsibilities for providing and installing items, such as miscellaneous steel used to stabilize
those structural members that support precast concrete
elements, must be clearly indicated in the contract bid
documents. The stability of a completed structure for
lateral loads is typically the responsibility of the EOR.

The load path of these floor forces must be followed
through the structure and considered in the design of
other members in the building. Because of the stresses
that may develop due to volumetric changes, tensile
connections parallel to the span at the bottom surface
of both ends of a floor member should not be used.
Even when torsion is resisted by a couple in the floor
elements in the completed structure, twisting of the
spandrel during erection prior to completion of connections must be considered. Spandrels that are pocketed to receive stems of the double-tee floor or roof

Fig. 4.2.9 Loadbearing spandrels.

Window

Spandrel
Wt.
Cast-in-place topping
e

W
Precast concrete floor

4.2.7 D
 esign Considerations for
Loadbearing Spandrels
Loadbearing spandrels support floor or roof loads.
Except for the magnitude and location of these additional loads, the design considerations are the same as
those for non-loadbearing spandrels.
Loadbearing spandrels support structural loads that
are usually applied eccentrically with respect to the
support. A typical arrangement of spandrel and supported floor is shown in Fig. 4.2.9. Loadbearing members loaded non-symmetrically may be subject to both
internal and external torsion. If the resulting applied
load is not coincident with the member’s shear center,
torsion will exist along the span of the member. Torsion
due to eccentricity must be resisted by the spandrel.
Potential rotation due to eccentricity is usually resisted by a horizontal couple developed in the floor
construction or by a coupled connection to supporting
columns. In order to prevent rotation of the spandrel,
the details must provide for a compressive force transfer at the top of the floor and a tensile force transfer
at the bearing of the precast concrete floor element.

4

Window

Window

Spandrel
Wt.
Cast-in-place topping
W
Precast concrete floor

e

Window

ARCHITECTURAL PRECAST CONCRETE

|

285

4

DESIGN

4.2.7 Design Considerations for Loadbearing Spandrels / 4.2.8 Design Considerations for Stacking Non-Loadbearing Panels

slabs decrease torsion stresses greatly, as well as minimize twist and eccentricity during erection.
If torsion cannot be accommodated by floor connections, the spandrel panel should be designed for the
induced stresses. Non-prestressed reinforced concrete
members subject to torsion should be designed in
accordance with the applicable provisions of the PCI
Design Handbook (MNL-120) and ACI Building Code
(ACI 318), Chapter 11. Prestressed members subject
to torsion should be designed in accordance with the
applicable provisions of the PCI Design Handbook and
ACI Building Code, recognizing that the member is
simply supported and capable of dissipating torsion effects into the diaphragm and connections.
Many precasters will choose to provide additional
connections along the length of a spandrel to prevent
cracking under anticipated gravity and lateral loads.
Adding unnecessary tiebacks is not advisable due to
the resulting build-up of restraint forces.
A loadbearing spandrel could be connected to a floor
or roof diaphragm (part of the lateral-load-resisting system) in more than one way. Structural integrity is typically achieved by connecting the spandrels into all or a
portion of the deck members forming the diaphragm,
which in turn would be connected to the supporting
beams and the beams would be connected to their
supporting columns. Alternatively, the spandrel could
be connected only to its supporting columns, which in
turn must then be connected to the diaphragm.

4.2.8 Design Considerations for Stacking
Non-Loadbearing Panels
Architectural precast concrete cladding panels are
usually independently supported. That is, each panel
has its own set of gravity and lateral connections to secure it to the building’s structural frame. Gravity load is
supported by the building columns and/or beams and,
thus, is transferred to the foundation. In most buildings, it is usually preferable to support panels in the
previously described manner.
However, there are some building types where it is
beneficial to take advantage of concrete’s inherent
strength and make the architectural concrete cladding self supporting. Only lateral tieback connections
are made to the building’s frame for lateral stability.
This design may provide economic benefits in structures where the exterior columns are set back from the

286

|

ARCHITECTURAL PRECAST CONCRETE

edge of the floor. One such building type is the suburban office building, which is usually a two- to six-story
steel-braced frame building. The exterior cladding consists of horizontal spandrels and vertical column covers
that are very repetitive in size, shape, and layout. The
vertical column covers can be designed to support the
weight of the spandrels, glass, and any other gravity
loads being carried by the stacked spandrels.
There are several advantages to stacking precast concrete cladding panels to support multiple stories. The
building frame can be lighter and less expensive because it resists the wind or seismic loads and does not
have to carry the weight of the precast concrete panels. Bearing brackets welded to the structure that are
required to individually support the panels are expensive; these are eliminated when panels are stacked.
There are some precautions to take when using the
stacking method. In the vertical direction, all of the vertical thermal volume change movements accumulate
at the top of the building. The designer must make
sure that precast concrete joint sizes and building
details allow for this without causing detrimental effects. The horizontal spandrel panels should hang on
the side of the column covers and not bear on top of
them (Fig. 4.2.10). This allows the spandrels to move
freely and accommodate axial volume change forces.
Because it is generally not desirable to grout joints between architectural panels, the gravity loads should
be transferred through shims. Shim sizes must be of
the appropriate size to keep bearing pressures to an
acceptable level. The joint at the bottom of column
covers should, however, be grouted if bearing pressures are high. Cladding panels should not be stacked
in high seismic risk areas where large in-plane building movements can occur (the design must allow for
building drift). Braced frames in low seismic risk areas
typically do not have appreciable lateral drift, making
stacking panels acceptable.
In low seismic risk areas, horizontal deflection of the
superstructure frame is only of moderate concern because the steel frame can drift and the simple-span
behavior of the panels is preserved. The critical detail is
the behavior of the joints in the corner. The wall parallel
to the direction of movement does not move, whereas
the wall perpendicular to the movement is dragged
along by the frame movement. Thus, story drift limits
in the range of height divided by 100 are appropriate
with 10-year wind loads. If the precast concrete walls

DESIGN

4.2.8 Design Considerations for Stacking Non-Loadbearing Panels / 4.2.9 Dimensioning Of Precast Concrete Units

Fig. 4.2.10 Stacked panels.

Spandrel

Column cover

Bearing pads
or drypack
Foundation

are buried (in lieu of a foundation wall), drift must be
limited to control cracking because these panels are
rotationally restrained at their bases.

4.2.9 D
 imensioning Of Precast
Concrete Units
As a general rule, precast concrete panels should
be made as large as possible without creating special
handling requirements or losing repetition. Flat panels should not be made any thicker than necessary
for economical and/or weight reasons. Neither should
they be made so thin that structural or performance
requirements cannot be fulfilled. Because of the multitude of combinations of sizes, functions, applications,
and finishes, no chart for sizing has been attempted.
However, here are some considerations for optimum
dimensioning.
The panel cross-section is generally chosen for architectural or aesthetic reasons. A panel that is too thin
may bow or deflect excessively, thereby creating caulking problems at the building corners or fit-up and leakage problems at attached windows, (see Table 4.6.1,
page 350).

4

The smaller the unit, the greater the number of pieces required for enclosure. More pieces usually means
more handling, more fastening points, and higher erection costs. Therefore, large units are preferable unless
they lack adequate repetition, which increases forming
costs or incur significant cost premiums for transportation. The widths of the panels are usually dictated by
architectural considerations or the structural grid of
the building frame. The maximum size of individual
units requires consideration of production repetition,
handling ease, shipping equipment required, erection
crane capacity, and loads imposed on a support system.
When desired, the scale of large panels may be reduced
by using reveals or rustication joints (Fig. 4.2.11).
Panels with a height of two or three stories are used
for low-rise buildings because of the simplicity of the
construction system. This may result in larger structural
sections in order to limit handling stresses during manufacture. For high-rise buildings it may be more expedient to work with story-height panels; the panels can
be more slender and the erection more simple.
Structural considerations, such as in-service loads,
will rarely govern, unless the panels are loadbearing
or where special conditions require a large spacing between connections. For most precast concrete exterior
bearing wall structures, except for tall, slender panels, the gravity dead and live load condition will control panel dimensions rather than load combinations,
which include lateral loads. Minimum dimensions for
grouting panels together at horizontal joints, space for
placing reinforcement, locating handling devices, or
accommodating a variety of connection conditions can
determine the minimum dimensions.
Reasonable slenderness ratios—minimum thickness
over unsupported length (the least distance between
members that provide lateral support [connections],
when in final position)—for flat panels should be:
• 1/20 to 1/50 for panels that are not prestressed.
• 1/30 to 1/60 for panels that are prestressed.
Higher ratios are feasible, but must be modified to
account for panel end fixity, deflection limitations, or
applied loads. However, all but the lowest slenderness
ratios should be subject to a careful structural and
performance analysis. For example, a more thorough
analysis is required in those buildings without effective
shearwalls for lateral loads or when walls are subjected
to large bending moments. Other types of walls, such
as window walls, should be individually designed.

ARCHITECTURAL PRECAST CONCRETE

|

287

4

DESIGN

4.2.9 Dimensioning Of Precast Concrete Units

Fig. 4.2.11
Albert G. Hill Building, One Hampshire Street, Cambridge, Massachusetts;
Architect: Ellenzweig, Moore and Associates;
Photo: ©1985 Steve Rosenthal.

288

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.2.9 Dimensioning Of Precast Concrete Units

For large, slender flat panels, the possibility of improving
structural capacities and overall performance by prestressing should be considered. The 7 in. thick (175 mm), 35ft
long column covers in Fig. 4.2.12 were prestressed because it was more economical than mild reinforcement.
Handling loads at the plant and during erection may
be minimized by special handling techniques, such
as tilt tables, vacuum lifting, cradles, or special lifting
frames. The precaster will normally endeavor to use
these methods rather than thicken the panels. The
shape of the member must provide (1) proper thickness
of concrete to develop insert capacities for handling
and for connections, (2) adequate cover of reinforcement, (3) ease of production, and (4) proper thickness
for the aggregate size used.
Practical considerations may govern thickness in order
to provide concrete cover and accommodate aggregate
sizes. Many surface treatments such as retarded finishes will reduce the cover and influence the minimum
thickness. Attention must also be given to scoring or
false joints in flat panels, where the required minimum
dimension should be measured from the back of the
panel to the bottom of the groove. Panel thicknesses

4

less than 4 in. (100 mm) may create problems for the
installation and proper concrete cover of lifting and
connection anchoring devices. Therefore, a 4 in. (100
mm) thick panel is the practical minimum thickness.
Performance requirements, such as sound attenuation, fire-rating, and the desired tolerances for planeness of façades, may govern minimum thicknesses of
flat panels. Sound insulation requires a sufficient mass
of concrete, which may govern thickness unless additional wall features help to dampen sound.
Rather than increasing the overall thickness of panels,
consideration may also be given to ribbing the panels. Ribs may be part of the architectural expression
or, where flat exposed surfaces are required, ribs may
be added to the back of panels for additional stiffness.
Stiffening ribs on the back may also be used as corbels
to transfer vertical loads to the structure.
Sculpturing of precast concrete units may increase
their structural strength. Such sculpturing may increase
the depth-to-span ratio by providing ribs or projections
in either direction of a unit. Contrary to common belief, reasonable sculpturing of a precast concrete cladding unit will not constitute a cost premium where

Fig. 4.2.12
Corporate Center, Franklin, Tennessee;
Architect: Little Diversified Architectural Consulting Inc.;
Photo: Marshall Bassett.

ARCHITECTURAL PRECAST CONCRETE

|

289

4

DESIGN

4.2.9 Dimensioning Of Precast Concrete Units / 4.2.10 Handling and Erection Considerations

sufficient repetition of the unit will keep mold costs
within reason and where the sculpturing will aid the
unit’s structural capacity. For further details refer to
Section 3.3.4.

Fig. 4.2.13 Optimum handling
sequence of precast concrete units.

1. Strip

2. Rotation to vertical
transfer to storage

3. Storage
in yard

4. Transport
to site

5. Erected
on building

Fig. 4.2.14 Hoisting and rotating multi-story units.

Rolling blocks

First step – Rigging onto unit on truck

Second step – Rotating

Third step – Unit
oriented for
installation

290

|

ARCHITECTURAL PRECAST CONCRETE

4.2.10 Handling and Erection
Considerations
Design consideration must be given to installation
conditions for the project with respect to handling,
transportation, and hoisting. This must be done before
sizes, shapes, and other features are finalized because
erection equipment will frequently influence panel
size. Preplanning for the fewest, quickest, and safest
possible operations that should be performed before
releasing the crane will greatly increase efficiency of
erection.
Transportation limitations have already been described in Section 3.3.10. Information on erection
procedures can be found in PCI Erectors’ Manual–
Standards and Guidelines for the Erection of Precast
Concrete Products (MNL-127).
The precaster is generally responsible for designing
panels for handling stresses and for the design of the
handling inserts. Units should be handled in a manner
to avoid structural damage, detrimental cracking, or
aesthetic impairment.
The optimum solution for economical handling is
the ability to strip a unit from the mold and tilt it into
a vertical position similar to the position of the unit
in its final location on the building (Fig. 4.2.13). This
solution is not possible when units are several stories
high (see Sections 2.4, 2.5, and 2.6). Such units may
be stored and shipped on their long side and handled
at the jobsite by rotating the units in the air with rolling
blocks (Fig. 4.2.14).
An important consideration in designing toward optimum erection costs—a significant portion of total installed cost—is to provide suitable access for trucks and
mobile equipment at the jobsite and sufficient room
and proximity to the structure to allow erection to proceed. Suitable access requires level, stable, and wellcompacted and maintained roadways or approaches
and consideration for snow removal. At the pre-job
conference, arrangements for jobsite access, onsite
storage areas, and other items affecting transportation
should be made with the GC or CM. Site conditions
should allow erection and transportation equipment to

DESIGN

4.2.10 Handling and Erection Considerations

proceed under their own power to a location (directly
adjacent to the building) where precast concrete members can be handled by the erection equipment directly
from the transportation equipment onto the structure.
Configuration of the building or adjacent buildings and
underground services or spaces may prevent erection
equipment from operating efficiently and may, in some
cases, require double handling of the units. The constraints of the site may determine the maximum size
and weight of panels that can be lifted into position.
It should be noted that cranes are rated by the safe
capacity they can lift with the shortest boom and at
the steepest boom-up angle. Maximum lifts for a given capacity crane will reduce rapidly as boom length
increases and the angle changes, and the designer
should, if possible, consider this during the planning
stage. For example, it is not advisable to design heavy
units for a high-rise building extending from a parking
or shopping plaza unless truck and equipment access
to the tower is ensured, at least until the units have
been installed on the tower, (Fig. 4.2.15). Erecting a
panel set back on a roof, such as a mechanical-equipment screen, may be problematic. As the crane’s boom
lays down to reach over a parapet, its capacity quickly
decreases. The entire job could be penalized by these
few difficult lifts, which could have been avoided had
the erection methods been considered. Early coordination of design and realistic erection plans are essential
for the overall economy of precast concrete.
To avoid unrealistic or impossible erection conditions, a review of the project by the designer with the
precaster and erector is beneficial. The erector should
state the requirements for handling and erecting the
members in the safest and most economical manner.
The designer should give careful consideration to all
the factors that affect the methods that can be used in
the construction of the project and prepare the design
to best suit jobsite conditions.
The precasters, in preparing their design, must first
determine where the lifting devices should be placed.
Lifting points should be compatible with the method
of shipping (flat or on edge) and placed so that the rigging does not interfere with the structural frame. The
same lifting devices should be used on the various precast concrete units so that frequent rigging changes
can be avoided. Only one type and size of lifting device
should be used in a precast concrete unit (bolted devices should not be used at one end and lifting loops

4

Fig. 4.2.15 Access for trucks and hoisting equipment.

DON’T
Double handling plus shoring
of roof and floors

DO
Erect panels on tower before necessary
access is blocked by lower building

DO
Reduce panel weight to suit
available hoisting equipment

at the other end of the unit). One set of lifting devices is required for handling and erection, and another
may be needed for stripping. Lifting devices should be
designed for actual loads plus an impact factor that
depends on the panel configurations and finishes.
Locating all lifting devices on the erection drawings allows a check for interference with other functions or
with finishing requirements.
Erection inserts cast in precast concrete members vary
depending on the member’s use in the structure, the
size and shape of the member, and the precast concrete manufacturer’s or erector’s preference.
Ideally the unit should be moved into position on the
building without having to be pulled back at the top or
the bottom of the unit. On occasion, the building configuration or the precast concrete unit may be shaped
so that a special hoisting or setting jig must be con-

ARCHITECTURAL PRECAST CONCRETE

|

291

4

DESIGN

4.2.10 Handling and Erection Considerations

Fig. 4.2.16(b)

Fig. 4.2.16(a)

structed to hold or “cradle” the unit for setting (Fig.
4.2.16a). These devices have been employed with success and economy where conditions justify their cost.
All temporary lifting and handling devices cast into
the members should be dealt with as specified on the
erection drawings based on one or more of the following conditions:
1. R
 emoved where they interfere with other trades.
2. R
 emoved and surfaces patched where demanded
by appearance requirements.
3. R
 emoved or protected from possible corrosion and
marring of the finished product where not exposed
to view.
4. N
 o action required.
If possible, the placing of lifting and handling devices
should be planned so that little or no patching will be
required after use. However, when the lifting and han-

292

|

ARCHITECTURAL PRECAST CONCRETE

dling devices are located in finished edges or exposed
surfaces, bolt or insert holes require filling and patching. Often, these devices may be recessed, filled, and
finished at a later date with prior designer approval.
However, specialized lifting equipment may be necessary to eliminate the necessity of patching exposed lifting and handling devices. A spreader bar with C-hooks
can be used to erect architectural precast concrete
panels without using exposed inserts (Fig. 4.2.16b).
Depending on local practice, the assignment of responsibility for the erection of precast concrete generally varies in the following manner:
1. The precaster is contractually responsible to the
GC and/or owner for the supply and erection of
the precast concrete products. Erection may be
performed by in-house labor or by an independent
erector hired by the precaster.

DESIGN

4.2.10 Handling and Erection Considerations

2. The precaster is responsible only for the supply of
the precast concrete product, F.O.B. plant, or jobsite. Erection is performed by either the GC/CM or
by an independent erector under a separate agreement with the GM/CM and/or owner.
3. The precaster is responsible for the supply of the
product to an independent erector who has a contractual obligation with the GM/CM and/or owner
to furnish and erect the complete precast concrete
structure.
On projects where precasters do not have the contractual obligation for erection, it is beneficial to the
project if they assign representatives to observe, coordinate, and report on erection activities. It is recommended that the precaster maintain sufficient contact
with the firm(s) responsible for both transportation
and erection to ensure that the precast concrete units
are timely delivered, properly handled, and erected according to design and project specifications. Erection
of architectural precast concrete should only be performed by qualified erectors employing workers who
are properly trained to safely handle and install the
product. Safety procedures for the erection of precast
concrete members are the responsibility of the erector,
and must be in accordance with all rules and regulations of local, state (province), or federal agencies that
have jurisdiction in the area where the work is to be
performed.
Proper planning of the construction process is essential for efficient and safe erection. The sequence of
erection must be established early and the effects accounted for in the bracing analysis and the preparation
of shop drawings.

4

ware and procedures, and instructions on removal.
Shoring and bracing should be installed as designed.
Units should never be left in an unsafe condition at
any time. Removal of temporary bracing and shores
is not permitted until proper alignment and adequate
permanent support has been provided or until it has
been authorized by the EOR to do so.
After erection, each panel must be stable and offer
resistance to wind, accidental impact, and loads that
may be imposed due to other construction operations.
Seismic forces during erection are usually only considered when the time of unsupported erection condition
will be unduly prolonged or when dislodgement could
cause progressive collapse. Inserts for attaching bracing are normally cast into the backs of the panels, but
may be field-drilled. Temporary bracing should be arranged so as not to interfere with other members being erected and other construction processes, and the
bracing must be maintained until permanent connections are accomplished. The single-story loadbearing
panels shown in Fig. 4.2.17 are temporarily braced until the connections are welded and floor or roof structural elements are installed. It is desirable that each
precast concrete element be braced independently of
other elements, so that an element may be moved,
if necessary, without affecting the adjacent element.
When possible, the final connections should be used
to provide at least part of the erection bracing, but
additional bracing apparatus is sometimes required
to resist all of the temporary loads. (Note: Only one
brace per panel is shown as the panel has already been
welded to adjacent panel.)

Generally, procedures to ensure stability during erection, such as bracing or temporary connection details,
should be developed by a licensed engineer or competent person engaged by the precaster or the erector.
Depending on the requirements of the local jurisdiction, the design sometimes requires review by the EOR
and/or building official.

The type of jobsite handling equipment selected may
influence the erection sequence, and, therefore, affect the temporary bracing requirements. Several types
of erection equipment are available, including truckmounted and crawler mobile cranes, hydraulic cranes,
tower cranes, monorail systems, derricks, and others.
PCI Erectors’ Manual (MNL-127) provides more information on the uses of each.

Erection drawings define the procedure on how to assemble the components into the final structure. When
required, the erection drawings should also address
the stability of the structure during construction and
any temporary connections required. When temporary
bracing, guying, or shoring is required, additional bracing drawings are recommended. These should show
items such as specific erection sequence, bracing hard-

Mobile cranes are most commonly used on buildings
up to about six-stories high, where access is adequate
and reasonable. Mobile cranes can handle larger panels than tower cranes. Monorail systems, fed by booms
and hoist or by cranes, are sometimes used by erectors
for buildings above 16 or 20 stories and where relocation of rails can be made in jumps of at least 10 to 15
stories. The use of either of these hoisting systems can

ARCHITECTURAL PRECAST CONCRETE

|

293

4

DESIGN

4.2.10 Handling and Erection Considerations

Fig. 4.2.17

operate independent of other trades requiring hoisting
operations.
When tower cranes are used for setting precast concrete units, their reach and capacity may determine
the panel maximum weight. Tower cranes, require
more than the usual planning because their structures,
foundations, and presence on the site are generally
permanent for as long as heavy construction phases
are on-going. The use of tower cranes may have a significant effect on the planning of the structural frame
and the sequencing of construction, which should be
considered during the planning stages of the structural
frame. Scheduling problems may be encountered unless firm time allocations for separate trades can be
maintained by the general contractor. Some erectors occasionally make use of nighttime hoisting with
tower cranes, transferring units to block and tackle
equipment or setting them temporarily for final placement or alignment during daylight hours. Stiff-legs or
elevators, combined with power buggies for transport
across floors, and many other handling innovations
have been developed for specific jobs.

294

|

ARCHITECTURAL PRECAST CONCRETE

Consideration should be given to the location of project man hoist(s) and material hoist(s) on high-rise towers. If the hoist(s) is located where precast concrete panels are to be placed, these panels will have to be lifted
off of the building and erected later with a large, expensive mobile crane with sufficient reach to the top of the
building. Alternately, pre-planning the location of the
hoist(s) at a window-wall location will minimize or eliminate expensive panel erection at the hoist bay area.
Hoisting and setting the precast concrete units are
usually the most expensive and time consuming processes of erection. Speed of erection is directly related
to the type of connections selected and the arrangement of the building frame. It is highly desirable that
connections allow for initial setting of the panel and
immediate release of crane, with final alignment completed independent of crane support.
Structural limitations governing the erection and/or
bracing sequence should be stated on the contract
drawings or in the specifications. Limitations may state,
for example, that loading of the structure shall be bal-

DESIGN

4.2.10 Handling and Erection Considerations / 4.2.10.1 Wall panels

anced, requiring that no elevation be erected more
than a stated number of floors ahead of the remaining
elevations; or limitations may involve the rigidity of the
structure, requiring that walls should not be erected
prior to completion of floors designed to carry lateral
loads. In steel frames, it should be determined how far
ahead final connections of the frame must be completed prior to panel erection. In concrete frames, it must
be determined what concrete strength is required prior
to imposing loads of the precast concrete panels.
The EOR should also recognize that connections between panel and frame impose concentrated loads on
the frame and that these loads may require supplementary local reinforcement. In the case of multistory
concrete frames, consideration should be given to the
effects of frame shortening due to shrinkage and creep.
Delay in erecting precast concrete panels to permit a
portion of the shrinkage and creep to occur may be
beneficial. For panels that are supported on a continuous bed of grout (such as in loadbearing wall construction), the maximum number of floor levels that can be
erected using only shims should be determined and, if
critical, indicated on the contract drawings. Where an
erection analysis is not performed by the EOR, or when
it is not the responsibility of the EOR, the contract documents should name the responsible party.

4

ing equal on the structure. It may be more economical
to minimize crane movement and finish an elevation as
much as possible prior to moving the crane.
Fig. 4.2.18(a - d) Typical rigging arrangements.
(a)

(b)

(c)

Erection procedures for precast concrete members
will vary depending on the size, shape, and design of
the members, the structural elements that will receive
or support them, and the overall complexity of the
structure. Typical rigging arrangements are shown in
Fig. 4.2.18 for architectural precast concrete panels.
Figure 4.2.18(a) shows the use of one rolling block,
while (b) shows the use of two rolling blocks. A spreader beam is shown in (c) and the use of an equalizing
rolling block is shown in (d).

(d)

4.2.10.1 W
 all panels
Panels should be shipped, whenever practical, to the
jobsite in a position so that turning is not required.
Wall panels should be rigged and hoisted onto the
structure near their final location and held in place until
safely secured. Final alignment and gravity and lateral
connections may be made at once or later by another
follow-up crew, depending on the type of connection.
The connections should be designed to allow for easy
adjustment in all directions. Panels should be installed
on a floor-by-floor basis where practical, to keep load-

ARCHITECTURAL PRECAST CONCRETE

|

295

4

DESIGN

4.2.10.1 Wall panels / 4.2.10.2 Columns

post-tensioning unless sufficient shims are provided.
The grout should achieve its required strength prior
to post-tensioning. For panels that are supported on
a continuous bed of grout the maximum number of
floor levels that can be erected using only shims should
be determined by the precast concrete design engineer
and, if critical, indicated on the erection drawings.
Structural stability of loadbearing panels is normally
achieved through connections to floor and roof diaphragms with high-capacity connections at stair walls,
elevator shafts, or other shearwall locations. All bracing should remain in place until stability is achieved by
completing connections and diaphragm is in place.
Fig. 4.2.19 Balance beam.

When wall panels are to be set back into the face of
the building under an overhanging structure, the panels cannot be picked up from above. These erection
difficulties may often be overcome by proper planning
or the use of specialized equipment. Temporary openings in floors above, or suitable scheduling of other
trades, can alleviate such difficulties. Whenever possible, contract documents should facilitate such provisions. Special equipment, such as a balance beam (flying jib) with a movable or fixed counterweight, may be
used to erect the panels without causing damage or
defacing the exposed surface (Fig. 4.2.19).
Consideration during construction for the stability of
a loadbearing structure is required during planning for
erection. Erection stability can sometimes be built into
the design of loadbearing elements. It may be possible
to design the foundations and anchor bolts to withstand
the forces generated from wind and temporary forces
caused by construction procedures. If base connections
do not provide sufficient stability for columns and walls
to be left free-standing, temporary guying and/or bracing should be provided until final structural stability is
achieved. It is desirable to start erection from a laterally
stable area, such as a corner wall or stair tower.
Base connections for multistory loadbearing panels
are normally achieved by weld plates or grouted mechanical sleeves. The walls may also be post-tensioned
vertically using high-tensile strength bars that couple
at pockets at the base of the panels at each floor level.
The base of the panels should be grouted or drypacked
before the next level of floor slabs is erected. Panel
bases should be grouted or drypacked prior to vertical

296

|

ARCHITECTURAL PRECAST CONCRETE

4.2.10.2 Columns
Leveling nuts provide the simplest method for erecting columns because the full load of the column can
be let off the crane and the column plumbed by adjusting the level of the nuts. The column is braced and
then the crane is released. The base plate and double
nut method is high in hardware costs but provides tolerance for out-of-place anchor bolts as the base-plate
holes may be slotted or made oversize.
The preferred procedure for heavy loads is a column
base plate with a single shim pack centered under the
column. The shims take full load and the anchor bolts
and nuts serve to stabilize the column. An advantage is
that the columns can be loaded immediately.
As panels and floor slabs are erected, columns can
start to bow or lean due to eccentric loading. Checking
of column plumbness and providing adequate shoring
are particularly important when using pretopped tees.
The heavier load of the pretopped members increases
the tendency for columns to come out of plumb. To
maintain a plumb condition, columns usually have guy
cables and/or pipe braces that allow the erector to
plumb and stabilize the column prior to, during, and
after the installation of upper-level beams and floor
slabs.
If columns are eccentrically loaded, they are often
erected slightly out of plumb on the side opposite the
eccentric load so that the eccentric load will bring them
back to plumb when fully loaded. The amount of “outof-plumb” is determined by trial and error for each
project and type of column. The column size, height,
and design affect its deflection under load.

DESIGN

4.2.10.3 Spandrels / 4.2.10.6 Protection during erection

4.2.10.3 S pandrels

special erection equipment is usually required.

Precast concrete spandrels usually extend from column line to column line at the building perimeter. They
connect either at the columns or the perimeter beams.
Spandrels, whenever practical, should be shipped to
the jobsite in the vertical position. Shoring or bracing
may be required to assist in erection until connections
are made.

Precast concrete soffits are normally shipped to the
jobsite in the horizontal position. All of the erection
methods are costly, time consuming, and require a great
deal of pre-planning and preparation. Consideration
should be given to combining soffit and spandrel in a
single unit.

Alignment and all gravity and lateral connections
should be made final, or the spandrels should be safely
secured, prior to releasing the load and disconnecting
the rigging. Long spandrels may be tied back at the
center, if required by the design of the member, to minimize panel bowing. Spandrels may be required to be
installed on a floor-by-floor basis to keep equal loading
on the structure, although this approach may not be
the most economical. It generally is more economical
to minimize crane movement and finish an elevation,
as much as possible, prior to moving the crane.

4.2.10.6 Protection during erection

Spandrels should be aligned to predetermined offset
lines established for each floor level. Vertical dimensions between spandrels should be checked with a
story pole or similar device to ensure that opening size
is within allowable tolerance.

4.2.10.4 C olumn covers and mullions
Column covers are usually manufactured in singlestory units and extend either from floor to floor or between spandrels. However, units two or more stories
in height may be used. Column covers and mullions
are usually shipped in the horizontal position. Two lines
may have to be used to rotate the member.
Column covers and spandrels are often erected sequentially. Caution should be used so that one member does not stress another.
When spandrel panels or column covers and mullions
are interspaced with strips of windows to create a layered effect of glazing, precast concrete, and glazing,
the GC should work closely with the designer to arrive at interface details that are able to be built within
the sequence of construction, and embody all the elements required of exterior wall design.

4.2.10.5 S offits
Soffit units are normally erected under perimeter
beams to form an architectural closure. To achieve this,

4

Rainwater, or water from hoses used during the construction of the building, can cause discoloration of
exposed precast concrete by first washing across other
building materials (such as steel, concrete, or wood)
and then across the precast concrete. The GC should
provide and maintain temporary protection to prevent
damage or staining of exposed precast concrete during construction operations. Particular care should be
taken to avoid jobsite water washing over the precast
concrete. Dirt, mortar, plaster, grout, fireproofing, or
debris from concrete placement should not be permitted to remain on the precast concrete and should be
brushed or washed off immediately with clean water. If
required, the final cleaning of the precast concrete units
should take place only after all installation procedures,
including repairs and joint caulking, are completed.
Precast concrete units and adjacent materials, such
as glass and aluminum, should be protected from
damage by field welding or torch cutting operations.
Therefore, the sequencing of work of other trades
should be taken into consideration by the GC/CM to
prevent such damage. Non-combustible shields should
be used during these operations. To minimize staining, all loose slag and debris should be removed when
welding is complete. All welds and exposed or accessible steel anchorage devices should be painted with a
rust inhibitive primer, or in cases of galvanized plates,
a cold galvanized coating (zinc rich paint containing
95% zinc). Such protection should be applied immediately after cutting or welding.
Apart from emphasizing care by the precaster, erector,
GC, and other trades, the architect during actual construction has little direct control over potential staining,
but may add the following recommendation as part of
the specifications:
“All staining and damage caused by other than the
precaster (such as oil from cranes and compressor
lines, bitumen from roofing operations, or smearing by

ARCHITECTURAL PRECAST CONCRETE

|

297

4

DESIGN

4.2.10.6 Protection during erection / 4.3 Contract Documents

caulking or painting contractors) should be repaired by
the precaster or by qualified personnel using methods
approved by the precaster under the responsibility of
the general contractor.”
Such repairs cannot be part of the precaster’s contract,
and the precaster should be compensated for repairing
any damage caused by others following supply of the
precast concrete (or erection, if part of the contract).

4.3 C ONTRACT DOCUMENTS
The architect’s contract documents (drawings and
specifications) should define the scope and intent of
the work required. These documents should supply the
information described in Section 4.1.2.
These documents should enable the precaster to
design for the forces that the architectural precast
concrete units must resist. This and other sections of
Chapter 4 should be thoroughly studied so that those
parties preparing the contract drawings understand
the required scope of those documents.
Translation of the design concepts into contract
drawings is relatively simple for typical units. Particular
attention must be paid to situations where non-typical conditions are encountered. These conditions may
include outside and inside corners, intermediate roof
levels, non-typical floors (such as ground level or mechanical floors), and entrances. Such details should not
be overlooked. Contract drawings should not be open
to different interpretations. Confusion may increase
costs during bidding, production or erection.
If design modules and master mold concepts are not
maintained during detailing, additional cost factors
may be introduced. The designer is advised to maintain
liaison with local precast concrete manufacturers because their services may be particularly valuable during
the development of working drawings. The final and
exact dimensions of all precast concrete units should
exactly replicate the preliminary design, subject to an
assessment of feasibility of all salient details.
Information in the contract documents should be
sufficient to enable the precaster to produce the shop
drawings to identify the information needed for precast concrete erection (coordination) drawings and
translate these details into precast concrete product
and erection requirements.
The designer should give sufficient details or descriptions on the drawings to clearly indicate all exposed

298

|

ARCHITECTURAL PRECAST CONCRETE

surfaces of the units and their respective finishes.
This is particularly important for returns and interior
finishes. The finish may be identified on the drawings
by coded numbers, by different shadings of the areas
representing exposed surfaces, or by any other method
that remains readable after printing or possible reduction of the drawings. To simply thicken the exposed
surface lines has proven less than satisfactory unless
carefully executed.
Holes, within agreed size limitations, in architectural wall panels for other trades should be made by
the precaster at the manufacturing plant and consequently should be detailed and located on the contract
drawings.
The use of isometric sketches often makes it simpler
to visualize details, particularly in the case of non-typical conditions. By making such isometric sketches part
of the working drawings, the designer greatly facilitates the interpretation of the project requirements.
Precaster’s shop drawings (erection and production drawings) translate project contract documents
into usable information for accurate and efficient
manufacture, handling, and erection of the precast
concrete units. These drawings are prepared under
the control of the precaster and should be prepared
in general conformance to PCI Drafting Handbook–
Precast and Prestressed Concrete (MNL-119) and the
project specifications.
The first step in preparing the shop drawings necessary for a project is a thorough review of the contract
documents (plans and specifications) to determine all
the factors that can influence decisions regarding the
precast concrete. The goal of this analysis is to produce
standardization of precast concrete units, and note
any modifications required of precast concrete units,
connections, shop production techniques, handling
methods, and erection plans. Aside from the general
architectural shape requirements, the main factor in
establishing standardization of the units is the building frame and its relationship to the architectural units,
connection locations, tolerances, and clearances.
Erection drawings should include all precast concrete
member piece marks, completely dimensioned size
and shape of each member, note the location of each
member with respect to building lines and/or column
lines and finished floors, and provide the details and
locations of all connections from member to member
or member to structure. Anchor drawings, which show

DESIGN

4.3 Contract Documents

the location of hardware supplied by the precaster to
be placed by the GC, should be prepared by the precaster. Joints and openings between precast concrete
members and any other portion of the structure should
be identified. These drawings are not necessarily intended to show or describe procedures for building
stability during erection. When temporary bracing is
required, additional bracing drawings are recommended. These may show such items as sequence of erection, bracing hardware and procedures, and instructions on removal.
Some precasters prepare in-house shop (production)
tickets from shop drawings, listing schedules of precast concrete units. Others produce separate drawings of each individual and different unit from typical
master mold units. When erection drawings contain all
information sufficient for design approval, production
drawings (except for shape drawings) need not be submitted for approval, except in special cases. When, the
architect requests record prints, the number required
should be stated in the specifications.
Specifications should state the number of copies of
erection drawings required for approval. Electronic
submissions may be requested.
Generally, shop (erection) drawings are submitted to
the GC who, after checking them and making notations, submits them to the design team for checking
and review. The drawings are routed to the subcontractors who must then check and coordinate their related
work. Final approved shop drawings are returned to the
precaster by way of the GC. The GC is responsible for
the project schedule, tolerances and dimensions of the
building frame, and coordination of the precast concrete
work with the work of other trades. Timely review and
approval of shop drawings and other pertinent information submitted by the precaster is essential. Fabrication
should not commence until final approval or “approvedas-noted” has been received. If shape drawings are submitted separately, approval must be obtained to allow
fabrication of molds and tooling. Alternatively, shop
drawings may be approved initially for mold production
and subsequently for panel production.
Because mold production requires the greatest
amount of production lead time, the common goal of
both the architect and the precaster at the shop drawing stage is to expedite all the details regarding the size
and shape of the precast concrete panels.
If stone veneer is to be incorporated into the precast

4

concrete, coordination between the precaster and
stone veneer supplier is required. The manufacture of
stone veneer panels requires a long lead time in order
to maintain the schedule. It is a common practice to
submit a preliminary set of shop drawings for stone
sizes and details for approval prior to submitting final precast concrete erection drawings for approval.
This allows the stone fabricator adequate fabrication
time to ensure a steady flow of stone to the precaster.
Separate subcontracts and advance awards often occur in projects with stone veneer panels. It is suggested
that the precaster detail all precast concrete units to
the point where the veneer fabricator is able to incorporate details, sizes, and anchor holes for individual
stone pieces.
While these procedures may affect normal submission routines, it is not intended that responsibilities for
accuracy be transferred or reassigned. In other words,
the precaster is responsible for precast concrete details
and dimensions and the stone veneer fabricator should
be responsible for stone details and dimensions.
The architect reviews the precast concrete manufacturer’s erection drawings primarily for conformance
to the contract documents, then passes them along
to the EOR for review of conformance to the specified loads and connection locations. This allows the
engineer to confirm his or her own understanding of
the forces on the structure at the connection points
and the precaster’s understanding of the project requirements. Design details, connection locations, and
specified loads cannot be left to the discretion of the
precaster. It is especially important in cladding panels,
where for example, a spandrel might weigh 20 tons
(18t) and its weight distribution imposes torsion and
shear forces onto the frame.
Eccentricity of weight can cause deformation of
the structure. To install the cladding within the specified tolerances, the structure must be strong and stiff
enough to resist both gravity and lateral forces without
excessive deflection.
“Approved” and “Approved-As-Noted” shop
drawings normally mean that the GC has verified dimensions to be correct and final for the following:
overall building dimensions, column centerlines, floor
elevations, floor thicknesses, column and beam sizes,
foundation elevations, the location of mechanical
openings, and other items pertinent to architectural
precast concrete. It also means that the design team

ARCHITECTURAL PRECAST CONCRETE

|

299

4

DESIGN

4.3 Contract Documents / 4.4 Reinforcement

has reviewed the precast concrete submission for general conformance with the design requirements of
the contract documents. Approval does not extend
to means, methods, techniques, sequences, or procedures of construction, or to safety precautions and
programs incident thereto, unless specifically required
in the contract documents.
Shop drawings will provide the GC and other trades
with a means of checking interfacing with adjacent
building materials and the precaster’s interpretation
of the contract drawings. They reduce plant costs and
speed production by providing effective communications between the architect and the production/erection departments of the precast concrete manufacturer.
The correct transfer of information from the contract
documents is the responsibility of the precaster.

4.4 R
 EINFORCEMENT
4.4.1 G
 eneral
In designing architectural precast concrete panels,
it is desirable that there not be any discernible cracking. In some cases, cracking may be permitted but
the crack width must be limited (see Section 3.5.17).
When a reinforced concrete element is subjected to
tensile stresses (likely due to flexure), the amount and
location of reinforcing steel has a negligible effect on
member performance until a crack has developed. As
stresses increase, hairline cracks may develop and extend a distance into the element. If cracks are narrow,
the structural adequacy of the element will remain
unimpaired.
In members in which concrete stresses are less than
the allowable tensile stress during service, distributed
reinforcement is needed to control cracking that may
unintentionally occur during fabrication, handling, or
erection and also to provide ductility in the event of
an unexpected overloading. In members in which the
stresses are expected to be greater than the allowable
concrete tensile stress, conventional or prestressed
reinforcement is required for satisfactory service load
performance, adequate safety, and to meet aesthetic
requirements. Reinforcement may serve any one of
these purposes in architectural precast concrete.
The types of reinforcement used in architectural precast concrete wall panels include welded wire reinforcement, bar mats, deformed steel bars, and prestressing
and post-tensioning tendons. Secondary reinforcement

300

|

ARCHITECTURAL PRECAST CONCRETE

may include an epoxy coated carbon fiber grid. Random
fibers, which are sometimes used to control shrinkage
cracks, do not have well defined structural properties
and, therefore, cannot be used to replace structural
reinforcement. Non-prestressed reinforcement is normally tied or tack welded together into cages by the
precaster, using a template or jig when appropriate, unless the precast concrete unit is a simple flat panel (see
Section 4.4.6). The cage, whether made for the entire
casting or consisting of several sub-assemblies, must
have sufficient three-dimensional stability so that it can
be lifted from the jig and placed into the mold without permanent distortion. Also, the reinforcing cages
must be sufficiently rigid to prevent dislocation during
consolidation of the concrete in order to maintain the
required cover over the reinforcement. The size of reinforcing bars is often governed by concrete dimensions
and the required concrete cover over the steel.
Reinforcement, in addition to that needed for structural reasons, is required to control thermal movement
and shrinkage, and the cracking that these might otherwise cause. As a general rule, bar sizes should be
kept reasonably small (No. 3 thru 6) even where this
will reduce the spacing of the bars (increase the number of bars). Smaller bars closely spaced will decrease
the size of potential cracks and improve distribution of
temperature stresses better than fewer, larger bars that
are widely spaced. The use of additional smaller diameter reinforcing bars as compared with fewer, heavier
bars becomes more important in thin concrete sections.
Because the sum of the widths of potential cracks in
concrete is more or less constant for a given set of conditions, the more well distributed cracks there are, the
smaller and less visible they will be.
In non-prestressed applications, the recommended maximum spacing of reinforcement in wall panel
wythes exposed to the environment is three times the
wythe thickness (18 in. [460 mm] maximum) for reinforcing bars, while common spacing for welded wire
reinforcement is 6 in. (150 mm). The size and spacing
will be function of strength requirements and crackcontrol criteria.
Large reinforcing bars create other problems, as well.
They often require anchorage lengths and hook sizes
that may be impractical, making supplemental mechanical anchorage necessary. Because reinforcing
bar termination points act as stress raisers and potentially cause shrinkage cracks, it is often better to use

DESIGN

4.4 Reinforcement / 4.4.2 Welded Wire Reinforcement

welded cross bars or other types of mechanical anchorage when the bars are large relative to the concrete
thickness, rather than rely on bond alone. Connection
details with reinforcing bars crossing each other require careful dimensional checking to ensure sufficient
cover.
Good bond between the reinforcing bar and the concrete is essential if the bar is to resist tension forces
and keep cracks small. Therefore, the reinforcing bar
surface must be free of materials that impair the bond
between the concrete and steel, including loose rust.
Mill scale that withstands hard-wire brushing or a coating of tight rust is not detrimental to bond.
The minimum reinforcement contained in each wythe
of non-prestressed members should be 0.001 of the
cross-sectional area of the wythe in each direction, except as otherwise required by analysis or experience.
Panels exposed to the environment that are less than
2 ft (0.6 m) wide in one direction may not need to be
reinforced in that direction — 4 ft (1.2 m) for panels
not exposed to the environment.
The reinforcing steel should be placed as symmetrically as possible about the panel’s cross-sectional centroid.
This is particularly true for flat and delicately shaped
panels. Non-symmetrical placement may cause panel
warpage due to restraint of drying shrinkage or temperature movements. Panels with a concrete thickness
of less than 8 in. (200 mm) may have reinforcement
placed in one layer; however, two layers of equivalent
weight are recommended to control concrete cracking
in thicker panels during handling. Strict quality control
is necessary to ensure adequate minimum cover is provided for all reinforcement, particularly in thin panels
and delicately shaped units. Suitable production techniques should be used to maintain this location during placement of the concrete. Reinforcing bars should
not be bent sharply around corners, especially in slender sections. Reinforcing bars should preferably terminate beyond an intersection of bars to ensure proper
anchorage.
Typical stirrups are not recommended for slender sections. They cannot always be bent accurately or to a
radius small enough to properly locate the main bars.
Bending difficulties and tight clearance tolerances
make the proper cover very difficult to achieve when
stirrups are used in wall panels. Precast concrete units
may contain attachment and lifting devices, as well as
prestressing steel anchorages, along with their associ-

4

ated reinforcement. The congestion that this additional
steel may cause, in conjunction with the normal reinforcement, should be given careful consideration.
If possible, reinforcing cages should be securely suspended from the back of the molds because spacers
may mar the finished surfaces of the panels. Metal
chairs, with or without coating, should not be used in
a finished face.

4.4.2 Welded Wire Reinforcement
Welded wire reinforcement (WWR) is the most common type of reinforcement used in architectural precast concrete. One or more layers of WWR may be
used as the main reinforcement or supplemented with
reinforcing bars in ribs and where otherwise required
to provide the area of steel required. WWR used in
architectural precast concrete is required to comply
with requirements of Building Code Requirements for
Structural Concrete (ACI 318). WWR is available in a
wide range of sizes and spacings, making it possible to
furnish the cross-sectional steel area required almost
exactly. Two sizes, a heavy and a light WWR, will usually suffice for most architectural precast concrete projects. The most common sizes of WWR used by precasters are 4 x 4–W4 x W4 (102 x 102–MW26 x MW26)
to 6 x 6–W2.9 x W2.9 (152 x 152–MW19 x MW19).
Different sizes may be considered standard depending
on geographic areas.
Welded wire reinforcement for architectural precast
concrete is supplied in flat sheets. Reinforcement from
rolls, if used in thin precast concrete sections, must be
flattened to the required tolerances. Many precast concrete plants have the capability of accurately bending
WWR to desired shapes, increasing its usefulness in large
members. Because the WWR is closely and uniformly
spaced, it is well suited to control cracking. Furthermore,
the welded intersections ensure that the reinforcement
will be effective close to the edge of the member, resisting cracking that may be caused by handling.
There is no specific ASTM specification for galvanized
WWR; however, ASTM A641, Zinc Coated (Galvanized)
Carbon Steel Wire is used as a reference. The amount
of zinc coating on wire reinforcement is rarely specified
for galvanized WWR. Galvanized wire can be produced
with thicknesses of zinc coating ranging from 0.30
oz/ft2 (107 g/m2) to 2.0 oz/ft2 (610 g/m2) for different
grades and wire sizes.

ARCHITECTURAL PRECAST CONCRETE

|

301

4

DESIGN

4.4.3 Reinforcing Bars / 4.4.4 Prestressing Steel

4.4.3 R
 einforcing Bars
Deformed mild-steel reinforcing bars are also used
extensively in architectural precast concrete. Deformed
reinforcing bars are hot-rolled from steels with varying
carbon content. Deformed bars conforming to ASTM
A615, Deformed and Plain Billet-Steel Bars for Concrete
Reinforcement, are generally available in No. 3 through
No. 11 in Grade 60 (420 MPa). Selection of grades of
reinforcing steel is determined by the structural design
of the precast concrete units. For bars that are to be
welded, ASTM A706, Low-Alloy Steel Deformed Bars
for Concrete Reinforcement, specifies a bar with controlled chemistry that is weldable with less preheat than
A615. Availability should be determined before ASTM
A706 bars are specified.

4.4.4 P
 restressing Steel
Prestressing may be used to minimize cracking of members by applying a precompression in the concrete that
counteracts the tensile stresses generated by the self
weight and applied loads. Prestressing may be either
pretensioning or post-tensioning. In either case, the prestressing force should generally be concentric with the effective cross-section. It is recommended that prestressing
in a panel, after all losses, be limited to the range of 150
to 600 psi. (1.0 to 4.1 MPa). A minimum of 225 psi (2.0
MPa) is required to be classified as a prestressed unit.
Prestressing in concrete panels may be used as partial
reinforcement for the following reasons:
1. Structural requirements for in-service loads.
2. Units are to be supported near the top and it is
desired to maintain the concrete in compression.
Units that are suspended should usually have light
welded wire reinforcing in both faces to counteract
the effects of tensile stresses. Alternatively these
units may be prestressed to provide a nominal residual compressive force.
3. Units are slender, and prestressing is chosen to facilitate handling without undue tensile stresses.
4. For general crack control, for example, handling
and erection of long (greater than 20 ft [6m]) insulated panels.
The belief that prestressing will reduce or control
bowing or warping is, in some cases, a misconception.
Unless tendons are located accurately (concentric with
the effective cross-section) and are securely maintained
in that location during casting, prestressing may actu-

302

|

ARCHITECTURAL PRECAST CONCRETE

ally increase the occurrence of bowing. Because panels
tend to bow outward due to thermal and shrinkage effects, and because outward bowing is generally more
objectionable than inward, some precasters choose to
force a slight inward initial bow by adjusting the location of the prestressing force.
Numerous wire, strand, and high-strength bar prestressing materials are available that are suitable for architectural precast concrete. Strands are commonly used in
both pretensioned and post-tensioned applications. They
are manufactured in two grades to conform to ASTM
A416, Uncoated Seven-Wire (Low-Relaxation) Strand
for Prestressed Concrete. High-strength, specially treated
bars used as prestressing tendons are covered by ASTM
A722, Uncoated High-Strength Steel Bars for Prestressing
Concrete. Uncoated, stress-relieved wires meeting the requirements of ASTM A421 can be used in post-tensioned
tendons. Unless prestressing is required for in-service load
or performance, the choice of using this type of reinforcement should be left to the precast concrete manufacturer.
A light bond coating of tight surface rust on prestressing tendons is permissible, provided strand surface shows no pits visible to the unaided eye after rust
is removed with a non-metallic pad.
In order to minimize the possibility of splitting cracks
in thin, pretensioned members, the strand diameter
should preferably not exceed the values in Table 4.4.1:
Table 4.4.1 Maximum strand diameter in wythe
thicknesses.

Wythe thickness,
in. (mm)

Strand diameter,
in. (mm)

less than 3 (75)

3

/8 (10)

3 (75) and thicker

1

/2 (13)

When exposed to view, tendon anchorages (stressing
pockets) should be recessed and packed with a minimum of a 1 in. (25 mm) thickness of low shrinkage,
non-metallic, concrete or grout and receive a sack finish. Prior to installing the pocket material, the inside
concrete surfaces of the pocket should be coated or
sprayed with an epoxy or latex bonding agent. This
seal should be adequately covered for curing because
shrinkage and contraction cracks allow access points
for moisture penetration and developing points for corrosion. When not exposed to view, the anchorage and
strand ends should be completely coated with a rust

DESIGN

4.4.4 Prestressing Steel

4

Fig. 4.4.1 Example of post-tensioned wall panel.
Extra layer of
mild reinforcement
at looped areas

2"
clear

A

1
/2" Plastic Coated
270 ksi strand (continuous)

3"

3"

32' – 0"

A

mild reinforcement

8' – 0"
8"

8"
SECTION A-A

8' – 0"
6"

Strand is extended beyond panel to allow length for the jack
3 x 6 in. pocket for strand anchor

inhibitor, such as bituminous paint, zinc rich paint, or
epoxy paint to avoid corrosion and possible rust spots.
Many architectural panels that do not lend themselves
to being pretensioned because of difficulties with long
line casting, such as jacking bulkhead or self-stressing
form requirements, can be easily post-tensioned. The
process of post-tensioning incorporates the installation
of either bonded or unbonded tendons in preformed
voids or ducts throughout the length of the member,
or through a section of the member. After the concrete has reached a predetermined strength, strands
are stressed and anchored against the hardened concrete. The flat panels in Fig. 4.4.1, the largest of which
measured 8 x 32 ft (2.4 x 9.8 m), were post-tensioned
in the precaster’s plant to provide optimum structural
integrity for panel lifting, hauling, and erection.
The strand (tendon) most frequently used in architectural precast, post-tensioned concrete is called the
monostrand and is composed of a single seven wire
prestressing strand. Although monostrands can be
fabricated to be grouted, they are usually coated with
grease or teflon and encased in a plastic tube. Thus,

Note: 1 in. = 25.4mm; 1 ksi = 6.895 MPa

they are typically used in the unbonded condition.
When panels are post-tensioned, care must be taken to ensure proper transfer of force at the anchorages. Provision for anchor plate and tendon protection
against long-term corrosion is essential. The anchorage
areas should be sealed immediately after the tendons
or strands are post-tensioned. Straight post-tensioning
strands or bars can be incorporated into the product,
and this would generally require anchorages at both
ends of the tendon. One method used to minimize
the number of anchorages is illustrated in Fig. 4.4.1.
Plastic-coated, unbonded strand with a low coefficient
of angular friction (μ=0.03 to 0.05) are looped within
the 8 x 32 ft (2.4 x 9.8 m) panels. Anchorages are installed at one panel end only.
A significant design consideration when determining
whether or not to employ prestressing is the evaluation of possible future unplanned openings. Cutting
of an unbonded tendon will remove the effect of prestressing for that particular tendon. While it is unlikely
that unplanned openings will be required, the designer
must still be cognizant of this.

ARCHITECTURAL PRECAST CONCRETE

|

303

4

DESIGN

4.4.5 Shadow Lines–Reflection Of Reinforcing Steel

4.4.5 S hadow Lines–Reflection
Of Reinforcing Steel
Reinforcing steel may show through some finishes
as light or dark shadow lines, usually directly over the
steel, depending on concrete mixture, vibration of reinforcement, or placing methods.
Steel reflection may be more likely to occur when
minimum cover is used over the steel and a rigid steel
cage is used. The cage will vibrate as an assembly in
phase (resonates) during the intense vibration used to
compact the low-slump concrete used in precast concrete units. Welded wire reinforcement or tack welding
of steel will stiffen the steel cage and is more likely to
produce reflection. Tied steel assemblies do not pose
this problem.
If shadowing does occur, sandblasting with the smallest available gun and nozzle and using fine grit can reduce the reinforcement outline to a reasonably uniform
tone. However, portions of the surface immediately
over the reinforcement may be less dense than areas
away from the reinforcement, making it extremely difficult not to over-erode the surface over the reinforcing
steel during sandblasting.
Another cause of steel reflection on the surface of a
panel is the use of galvanized reinforcement and/or the
use of galvanized and non-galvanized reinforcement
together in fresh concrete (unless the steel surface of
the galvanized reinforcement is passivated).
The reactions of zinc in a concentrated alkaline material (concrete prior to setting has a pH of 12.5 to 13.5)
liberate hydrogen gas, forming bubbles of gas at the
zinc-coated surface during initial stages of hydration.
These bubbles or voids may cause local porosity of the
concrete and increase the water absorption of the concrete over the steel, resulting in steel reflection.
When galvanized and non-galvanized reinforcement
or steel forms are in contact in fresh concrete, galvanic
cell problems may arise during the initial processes of
hydration. An electrochemical reaction occurs in which
zinc is consumed to form either zincate ion or zinc hydroxide on the anodic galvanized steel, and hydrogen
gas may be liberated on the cathodic non-galvanized
steel or locally on cathodic areas of the zinc. This reaction tends to be more active where there is a high
ratio of black steel surface area relative to galvanized
steel surface area. In this case, rows of hydrogen bubbles may form along non-galvanized reinforcement in

304

|

ARCHITECTURAL PRECAST CONCRETE

contact with galvanized reinforcement, or occasionally
on smooth black steel form surfaces, resulting in steel
reflection.
Although not a steel reflection problem, discoloration
or surface distress may occur with some form oils or
mold release agents, and concrete may stick in places
to the black steel form if galvanized steel is in electrical
contact with the form surface.
Research has shown that proper treatment of reinforcement can suppress the liberation of hydrogen on
a galvanized steel surface exposed to alkaline cement
paste. The following is an excerpt from ASTM A767,
Zinc-Coated (Galvanized) Steel Bars for Concrete
Reinforcement:
“4.3 Chromating — The galvanized coating shall
be chromate treated. This is to preclude a reaction
between the bars and the fresh portland cement
paste. Proprietary chromating solutions of equivalent
strength are permitted in place of the generic chemical treatment specified.
4.3.1 If the chromate treatment is performed immediately after galvanizing, it may be accomplished by
quenching the reinforcement bars in a solution containing at least 0.2 mass (weight) percent of sodium
dichromate in water (such as 2 kg/m3 [3 oz to each
10 gal] of quench water) or by quench chromating in
a minimum of 0.2% chromic acid solution. The solution shall be at least 32 °C (90 °F). The galvanized
reinforcement bars shall be immersed in the solution
for at least 20 seconds.
4.3.2 If the galvanized reinforcement bars are at
ambient temperature, the chromate treatment shall
be the same as specified in Section 4.3.1 except that
a 0.5 to 1.0% concentration of sulfuric acid shall be
added as an activator of the chromate solution. In
this case, there is no temperature requirement for the
activated chromate solution.”
Despite the above precautions, galvanizers are probably not using the required sodium dichromate because the Environmental Protection Agency (EPA) has
determined it is a potential carcinogen (cancer causing
substance). If the reinforcement was not chromated,
another solution would be to add chromium trioxide
(CrO3—chromic acid anhydride) to the concrete (150
parts per million based on weight of mixing water or
about 8 oz of a 10% solution per cubic yard of concrete). Alternatively, consider a chromate surface treatment (brushing galvanized steel with 0.3ONK2Cr2O7

DESIGN

4.4.5 Shadow Lines–Reflection Of Reinforcing Steel / 4.4.7 Corrosion Resistance of Reinforcement

+ 0.4ONHNO3 and then rinsing with water) that suppresses hydrogen liberation. Although these chromate
solutions may also be considered carcinogenic.
Portland cements in the United States contain an average of 52 ppm of CrO3 based on weight of mixing
water per cubic yard of concrete with a range of 10
to 156 ppm. Therefore, some cements may have adequate chromium to passivate the surface of the galvanized steel and prevent the problems.
If neither of these approaches is appropriate, it may
become necessary to age the shiny zinc film so that a
dull, matte, gray or whitish film appears on the surface.
This requires exposure of the bright zinc coating to water and sunlight. Weathered or aged galvanized steel
is less susceptible to the problems described because
when the zinc has a patina the zinc is less active.

4.4.6 T ack Welding
Tack welding, unless done in conformance with AWS
D1.4, may produce embrittlement or metallurgical
notch of the reinforcing steel in the area of the tack
weld. Tack welding seems to be particularly detrimental to ductility (impact resistance) and fatigue resistance. Tack welding affects static yield strength and
ultimate strength to a lesser extent. Where a small bar
is tack welded to a larger bar, the detrimental “metallurgical notch” effect is exaggerated in the large bar.
Fast cooling under cold weather conditions is likely to
aggravate these effects.
Welding procedures are critical for reinforcing steel
because this steel has a relatively high carbon content. The more carbon the steel contains, the more
brittle the material and the higher the susceptibility to
embrittlement occurring after a weld begins to cool.
For example, if a welded assembly that had not been
welded using proper procedures is raised to shoulder
height and dropped to the floor, it is quite possible the
bars would break off at the weld point like a shattered
piece of crystal. This is due to the brittle formation of
martensite.
AWS D1.4 prohibits the use of tack welds that do not
become part of the permanent welds of reinforcing
steel, unless approved by the precast concrete design
engineer. However, tack welding of reinforcement at
locations where neither bar has a structural function
should be allowed. For example, welding the ends of
the outside bars (within 10 bar diameters from the free

4

end of the bar) may be an aid in fabrication of reinforcing cages.
When bars are bent cold without the addition of heat,
they become sensitive to heat. Subsequently, the application of too much heat will cause mechanical property
changes in the cold-worked area of the bar and result
in unpredictable behavior of the reinforcing bar at the
bend. Therefore as a precaution, it is necessary to keep
welds away from cold bends. While AWS D1.4 suggests allowing a cold bend at two bar diameters from a
weld, experience shows that a minimum distance of 2
in. (50 mm) with 3 in. (75 mm) preferred is better with
the small bars commonly used in precasting.
Tack welding must be carried out without significantly diminishing the effective steel area or the bar area
should be one-third larger than required. A low heat
setting should be used to reduce the undercutting to
1
/16 in. (1.6 mm) of the effective steel area of the reinforcing bar.
The welding of crossing bars to assemble reinforcement may be done by fusion welding. A large number of tensile and bend tests by independent labs have
confirmed that controlled welding does not adversely
affect the mechanical properties of the bars. As a result of these findings, PCI endorses the use of fusion
welding in the fabrication shop, but still recommends
that field tack welding should not be permitted unless
authorized by the design team.

4.4.7 Corrosion Resistance of
Reinforcement
A key concern for owners and designers when planning a construction project comes from the possibility of
corrosion of reinforcement. Corroding reinforcements
can cause significant, long-term deterioration that may
go unnoticed until excessive damage has occurred.
However, techniques used to design and install reinforcement in all precast concrete applications inherently
prevents this from happening. Reinforcement surrounded by concrete is protected from corrosion by the alkaline nature of the concrete. This is particularly important
in architectural precast concrete because buildings with
such members need to retain their attractive appearance and durability over a long period of time.
In an alkaline environment, a very thin protective layer
of iron oxide (called the passivating layer) forms on the
steel surface-to-concrete interface, which stops corro-

ARCHITECTURAL PRECAST CONCRETE

|

305

4

DESIGN

4.4.7 Corrosion Resistance of Reinforcement / 4.4.7.2 Concrete Cover

sion from starting. Passivity may be destroyed either
by carbonation of the concrete, which reduces its alkalinity, or by the ingress of chlorides, which can cause
localized disruption of the passive layer.
Depassivation, either locally or generally, is necessary but not sufficient cause for active corrosion-induced damaged to occur. The presence of moisture
and oxygen are essential for corrosion to proceed at
any significant rate. Because architectural precast concrete members are vertical or inclined, the possibility of
moisture retention or ponding on these elements is remote. Therefore, such members are not as susceptible
to moisture penetration.

4.4.7.1 C hlorides
Chloride ions are most common cause of corrosion of
reinforcing steel in concrete. In general, precast concrete does not have a quantity of chloride ions that
will cause corrosion. This is the case even in marine
environments with salt spray, fog, or mist. The principal potential source of chloride in architectural precast
concrete is that introduced into the concrete, intentionally or otherwise. The use of chloride-containing
materials in concrete is strongly discouraged.
Architectural precast concrete is generally not exposed
to deicing salts. According to Table 4.4.1 of ACI 318,
the total mixture water-soluble chloride ion content
contributed from the water, aggregates, cementitious
materials and admixtures should not exceed 0.06%
chloride ions by weight of cement for prestressed concrete. The corresponding figure for reinforced concrete
is 0.15%. Additionally, each admixture should not contribute more than 5 parts per million by weight of chloride ions to the total concrete ingredients.
In situations, where chloride ions reach the steel,
some of them find naturally occurring imperfections
in the passivating layer causing the steel to initiate
corrosion.
The threshold value of chloride concentration from
externally applied sources, below which corrosion
does not occur, is about 0.20% acid-soluble chloride
ion content by weight of cement. This is equivalent
to about 0.025 to 0.040% by weight of concrete or
1.0 to 1.5 lb/yd3 (0.593 to 0.890 kg/m3) of concrete.
It can take years or decades for this threshold amount
of chloride to collect at the steel in uncracked concrete
from externally applied salts. The time to corrosion de-

306

|

ARCHITECTURAL PRECAST CONCRETE

pends on the concrete cover and the concrete permeability to chloride ions. Permeability is mainly affected by
the water-cement ratio (W/C) and concrete additives.
While chlorides are directly responsible for the initiation of corrosion, they play only an indirect role in
determining the rate of corrosion after initiation. The
primary rate controlling factors are the access of oxygen, the electrical conductivity, and the relative humidity — all of which are interrelated — and temperature.
Similarly, carbonation destroys the passive film but does
not play a role in determining the rate of corrosion.

4.4.7.2 Concrete Cover
Two key factors that influence the likelihood of reinforcement corrosion initiation in precast and prestressed concrete members are the amount of cover
for the reinforcement and the properties of the concrete, particularly the concrete’s permeability, immediately surrounding the reinforcement and whether it is
cracked. Ensuring proper cover can be an overlooked
item in a durability protection program. Economical
corrosion-resistant design depends on understanding
how these two factors influence corrosion.
Concrete cover refers to the minimum clear distance
from the reinforcement to the surface of the concrete.
For exposed aggregate surfaces, the concrete cover to
the steel’s surface should not be measured from the
original surface. Instead, the depth of mortar removed
from between the pieces of coarse aggregate (depth of
reveal) should be subtracted to give a realistic measurement. Attention also must be given to scoring, false joints
or rustications, and drips, as these reduce the effective
cover. The reduction in these areas should not exceed
one-third of the specified cover. The required minimum
cover should be measured from the thinnest location to
the reinforcement. In order to provide corrosion protection to reinforcement, concrete cover should conform to
ACI 318, Section 7.7, as listed in Table 4.4.2.
Cover requirements over reinforcement should be
increased to 11/2 in. (38 mm) for non-galvanized reinforcement or be 3/4 in. (19 mm) with galvanized or epoxy-coated reinforcement when the precast concrete
members are acid-treated, exposed to a corrosive environment, or subjected to other severe exposure conditions. For all exposure conditions, the cover should be
greater than one-and-a-half times the nominal maximum aggregate size or equal to the specified concrete
cover, whichever is larger. This minimum cover also al-

DESIGN

4.4.7.2 Concrete Cover / 4.4.7.4 Carbonation

Table 4.4.2. Minimum Cover Requirements for Architectural
Precast and Prestressed Concrete.1, 2

Condition

Minimum Cover3

Exposed to earth or
weather

#11 bar and smaller and
prestressing tendons
11/2 in. and smaller – 3/4 in.

Not exposed to
earth or weather

#11 bar and smaller – 5/8 in.;
prestressing tendons
11/2 in. and smaller – 3/4 in.

Note: 1 in. = 25.4 mm
1

Manufactured under plant-controlled conditions.

Increase cover by 50% if tensile stress of prestressed members exceeds 6 f′c .

2
3

 ealistic only if maximum aggregate size < 1/2 in. (13mm) and reinforcing
R
cage is not complex.

lows for proper flow and consolidation of the concrete
under the reinforcement as the concrete is placed.
Reinforcement should be placed within the allowable
tolerances, but the concrete cover should be set so
that the resulting concrete cover is never less than that
specified after consideration of maximum reinforcement placement tolerances. Where possible, excess
cover of 3/8 in. (10 mm) should be specified depending
on the degree of complexity of the cage because of the
tolerance of locating reinforcing steel using standard
fabrication accessories and placing procedures.
In determining cover, consideration should be given
to the following:
1. Maximum aggregate size, as cover should always
be greater than the nominal maximum aggregate
size, particularly if a face mixture is used.
2. Means used to secure the reinforcement in a controlled position and maintaining this control during
placement of concrete.
3. Accessibility for placement of concrete, and the
proportioning of the concrete mixture relative to
the environment.
4. Type of finish treatment of the concrete surface.
5. Service environment at the concrete surface: interior or exposed to weather, ocean atmosphere, or
corrosive industrial fumes.

4

cial, although ACI 318 Section 7.7 does not specifically
say so, and 21/2 to 3 in. (63 to 75 mm) should be regarded as a maximum. Concrete cover thickness outside this
limit lacks the restraint of the steel and is consequently
more likely to have wide cracks. In no case should the reinforcement be allowed to be beyond mid-depth of the
precast concrete element from the exposed surface.

4.4.7.3 Permeability
The ability of high-quality architectural precast concrete to resist the ingress of water, carbon dioxide,
chloride, oxygen, or other deleterious substances depends mostly on the permeability of the cement paste.
Because the aggregate particles are surrounded by the
hardened cement paste and most sound aggregates
have low porosity, the permeability of concrete is principally a function of the permeability of the cement
paste component of the concrete.
The water permeability of hardened cement paste is
primarily a function of the original w/c and the length
of the curing period (extent of hydration). Low permeability is obtained in a well-consolidated concrete with
a low w/c , a characteristic of architectural precast concrete. A maximum w/c of 0.40 is recommended for corrosion protection of concrete exposed to deicing salts,
brackish water, seawater, or spray from these sources.
Figure 4.4.2 shows how w/c affects the penetration of
chlorides in uncracked concrete. If the minimum concrete cover required by ACI 318, Section 7.7 is increased
by 1/2 in. (13 mm), the w/c may be increased to 0.45 for
normalweight concrete. Prior to specifying water-cement
ratios less than 0.40, the design team should contact
local precast concrete suppliers to determine their capabilities with the desired special-facing aggregates.
Permeability can also be reduced by the use of clear
penetrating concrete surface sealers, such as silane or
siloxane. These are chemical products that partially
penetrate the concrete and prevent moisture or soluble chemicals from penetrating to the reinforcement.
However, prior to specifying sealers, designers should
contact local precasters to determine the probable effect of specific sealers on surface appearance and project cost.

6. Fire code requirements.
Increased concrete cover improves the protection provided to the reinforcement, in part, because it acts as a
diffusion barrier to the passage of water vapor and liquid. Providing more cover than necessary is not benefi-

4.4.7.4 Carbonation
The reaction of carbon dioxide (CO2) and sulphur
dioxide with hydrated portland cement is called carbonation, which neutralizes the passivating oxide film

ARCHITECTURAL PRECAST CONCRETE

|

307

4

DESIGN

4.4.7.4 Carbonation

Fig. 4.4.2 Measured chloride profiles in moist-curved concrete from two FHWA investigations.

0.5

Average Chloride Ion after 830
daily salt applications (outdoor)

W/C = 0.51

Average Chloride Ion after 44
weeks (indoor)

Acid-Soluble Chloride Ion Content,
Percent by Mass of Concrete

0.4

Reference:
Sherman, M.R., McDondald, D.B., and Pfeifer,
D.W., “Durability Aspects of Precast Prestressed
Concrete — Part 1: Historical Review,” PCI
JOURNAL, V. 41, No. 4, July-August 1996, pp. 62-74

0.3

1 in. = 25.4 mm

0.2
W/C = 0.40

W/C = 0.50

Chloride Ion
Threshold

0.1

W/C = 0.60

W/C = 0.40
W/C = 0.28

0
0

1

2

3

Chloride Sample Depth, in.

on the steel developed by the alkaline conditions of
hydrated cement paste. Normally, with high-quality
concrete, the effect of carbonation does not penetrate
more than about 1/8 to 1/4 in. (3 to 6 mm), even when
exposed to the weather for 30 years. The effects of
carbonation can cause the pH of the concrete’s pore
water to drop from between 12.6 and 13.5 to about 8
to 9, at which level the passive film on steel is unstable.
Thus, if the entire concrete cover layer were carbonated, corrosion of steel would occur, provided oxygen
and moisture necessary for the reactions of corrosion
are present.
As a result, it is important to know the depth of carbonation and, specifically, whether the carbonation
“front” has reached the surface of the embedded steel.
Because of the presence of coarse aggregate, the front
does not advance as a perfectly straight line. It might
also be noted that, if cracks are present, CO2 can enter
through them so that the front advances locally from
the penetrated cracks. Galvanized steel remains passivated to much lower levels of pH than does black steel.
Carbonation rates are generally low in typical precast concrete with w/c less than 0.40, but are on the
rise because of the increased concentration of gases
in industrial environments. Under natural conditions,
the atmospheric concentration of CO2 in air is about

308

|

ARCHITECTURAL PRECAST CONCRETE

0.03%; in U.S. cities this is increased to 10 times that
value and, at industrial sites, can rise as high as 100
times normal levels.
Carbonation occurs progressively from the outside of
concrete exposed to CO2, but does so at a decreasing
rate because CO2 has to diffuse through the pore system, including the already carbonated surface zone of
concrete. Carbonation does not occur in very dry concrete or concrete at 100% RH. Apparently, at 100%
RH, the moisture blocks CO2 from passing through the
pores. The most aggressive environment for concrete
carbonation is that of alternate dry and wet cycles
(and, of course, high temperatures). Under constant
conditions, the optimum environment for carbonation
occurs at a RH of between 50 and 70%.
The considerable influence of concrete’s moisture content on carbonation means that even in a single building made of the same mixture design, there may be
considerable variation in the depth of carbonation at a
given age. Walls that are more exposed to rain will have
a lower depth of carbonation, as will sloping surfaces
that can be washed down by rain. This also applies to
walls that can be dried thoroughly by the sun’s rays.
The chloride content at the carbonation front reaches
much higher levels than in uncarbonated concrete as any

DESIGN

4.4.7.4 Carbonation / 4.4.7.5 Crack Widths

chloride bound in the cement paste is released and is
also much higher than the levels measured just below
the concrete surface. Added to this factor is the effect
of the decrease in pH of the carbonated concrete. Note
that the concentration of chlorides necessary to initiate
corrosion (the threshold value) decreases with pH.

4.4.7.5 C rack Widths
Cast-in-place concrete structures are inherently prone
to restrained volume change–induced cracking. This
cracking results from the normal restraint of movement that occurs when complex shapes and structural
elements attempt to contract during early-age, concrete volume changes. These changes are caused by
concrete temperature changes and drying shrinkage
and are normal for cast-in-place concrete. Such natural restraint is difficult to eliminate. Therefore, cast-inplace concrete is designed with the assumption of a
cracked section.
Precast concrete has less natural restraint problems,
because the individual members are not initially integral
with the structure. This is particularly true during the
early-age time period when large concrete-related thermal and drying shrinkage effects occur. Sufficient reinforcement must be used in each unit to control the distribution of any shrinkage cracking which may occur.
The potential exposure of the reinforcement to oxygen and moisture as a result of cracking (both necessary for corrosion) is lessened in precast concrete. In
addition, the denseness and impermeability of the precast concrete due to low water-cement ratios of plantcast concrete should be considered when evaluating
the corrosion potential.
A certain amount of cracking may occur without having any detrimental effect on the structural capacity of
the member; it is impractical to impose specifications
that prohibit cracking. However, in addition to being
unsightly, cracks are potential locations of concrete
deterioration and should be avoided if possible. The
key point is cracks do not always result in corrosion of
reinforcement. This depends not only on the width of
the crack and whether it reaches the reinforcing steel,
but also on the presence of chlorides or low pH in combination with oxygen and moisture.
When corrosion starts at a crack, the depassivated
areas near the crack become the anode of a corrosion

4

cell, while portions of the bar that are still protected by
sound, alkaline concrete become the cathode. At the
anode, metal ions are released. At the cathode, oxygen combines with water to form hydroxyl ions, which
flow through the electrolyte to the anode where they
combine with the metal ions to produce iron hydroxide. As a secondary reaction, this hydroxide combines
with additional oxygen to form rust. The problem is
rust occupies more volume than steel, causing more
cracking.
The rate at which corrosion will occur depends on
the resistance of the path through the concrete between the anode and the cathode and the availability
of oxygen at the cathode, which is situated on the reinforcing bar surrounded by sound concrete, not at the
crack. The rate of corrosion is thus dependent on the
properties of the sound concrete. Cracks with widths
below a certain limiting magnitude play no role in the
reinforcement corrosion process.
The amount and location of reinforcing steel has a
negligible effect on structural performance of a concrete element until a crack develops. As tension stresses increase above the concretes’ modulus of rupture,
hairline cracks will develop and extend a distance into
the member. A sufficient amount of closely spaced reinforcement limits crack widths and the intrusion of
corrosion initiators. Prestressing also may be used to
limit or eliminate crack width. If the crack width is narrow, not over 0.012 in. (0.30 mm), the structural adequacy of the casting will remain unimpaired and the
crack will have little influence on the potential for corrosion of the reinforcement. If crack repair is required
for the restoration of structural integrity, cracks may be
filled or pressure-injected with a low-viscosity epoxy.
Cracks transverse to the reinforcing steel usually do
not cause continuing corrosion of the reinforcement
if the concrete has low permeability. The potential
corroded length of intercepted bars is likely to be no
more than three bar diameters, because the exposed
portion of a bar at a crack acts as an anode. Cracks
that follow the line of a reinforcing bar are much more
damaging because the potential corroded length of
the bar is much greater and the resistance of the concrete to spalling is reduced. For surfaces exposed to
the weather, cracks up to 0.005 in. (0.13 mm) wide
have no influence on the corrosion of reinforcement
and should be acceptable from an aesthetic viewpoint
(Fig. 4.4.3).

ARCHITECTURAL PRECAST CONCRETE

|

309

4

DESIGN

4.4.7.5 Crack Widths / 4.4.7.7 Corrosion Protection

Fig. 4.4.3 Viewer reaction to cracks of different widths.
Crack Width (mm)
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

100
90

Architects
Engineers
Lay Public

80
79
Viewers’
Finding
Crack
Acceptable
(%)

60
50
40
30
20
10
0
0.005

0.010

0.015

0.020

0.025

0.030

0.035

Crack Width (in.)

Figure 4.4.4 shows variations in crack width between
the concrete surface and the bar surface obtained by
injecting resin into cracks in a beam while under load
and then sawing up the beam after the resin has set.
For a given steel stress, the crack width at the bar surface remains constant while the width at the surface
varies more or less linearly with the increase in cover.
There is not a constant relationship between surface
width and width near the bar. A direct relationship between surface crack width and corrosion should not be
expected to exist.

4.4.7.6 Corrosion Protection
Corrosion of reinforcement is usually not a problem
in architectural precast concrete. It should be recognized that galvanized or epoxy-coated reinforcement
cannot take the place of quality control in mixture proportioning and steel placement. Sound concrete having strengths of 5000 to 6000 psi (34.5 to 41.4 MPa)
in 28 days with a w/c of 0.40 or less and proper cover
(greater than 3/4 in. [19 mm]) typically will provide all of

Crack width appears to have a
significant effect on the amount of
corrosion at a relatively early age,
because the time to depassivation
is dependent on the width of the
crack. However, once corrosion
has started, the rate of corrosion is
independent of crack width. If the
combination of concrete density
and cover thickness is adequate
to restrict the flow of oxygen and
moisture, then the corrosion process is self-sealing. Crack widths
still should be controlled for appearance and watertightness.

310

|

ARCHITECTURAL PRECAST CONCRETE

Dis. from bar surface (mm)

Fig. 4.4.4 Internal crack widths.

Note: 1 mm = 0.039 in.; 1 N/mm2 = 145 psi

80

75 mm cover
(3 in.)

80

75 mm cover
(3 in.)

60

57 mm cover
(21/4 in.)

60

57 mm cover
(21/4 in.)

40

38 mm cover
(11/2 in.)

40

38 mm cover
(11/2 in.)

20

0
0

20

0.1
Crack width (mm)

(a) Steel stress = 138 N/mm2

0.2

0
0

0.1

0.2
0.3
Crack width (mm)

(b) Steel stress = 207 N/mm2

DESIGN

4.4.7.7 Corrosion Protection

the corrosion protection necessary for the reinforcing
steel in architectural precast concrete.
GALVANIZED REINFORCEMENT. The term galvanized steel refers to the electrolytic method that bonds
zinc to the steel surface. Zinc is a reactive metal that
readily oxidizes in air to form a corrosion-resistant film
of zinc oxide. The zinc oxide layer is very thin, hard,
and tenacious and is the first step in the development
of the protective corrosion product-layer normally associated with galvanized coating. When this surface
has access to freely moving air in normal atmospheric
exposure, the surface reacts with rainfall or dew to
form zinc hydroxide.
During drying, the zinc hydroxide reacts with carbon
dioxide in the atmosphere and is converted into a thin,
compact, and tightly adherent layer of basic zinc carbonate. This layer provides the barrier protection afforded by the galvanized coating. Because it is relatively
insoluble, the basic zinc carbonate layer is weather resistant and, once formed, minimizes further corrosion.
Galvanized WWR generally is manufactured from galvanized wire and is a stock item in some sizes. There
is no specific ASTM specification for galvanized WWR,
but ASTM A641/A641M serves as a reference. The
amount of zinc coating on WWR is rarely specified for
galvanized WWR. Galvanized wire can be produced
with thicknesses of zinc coating ranging from 0.30 to
2.0 oz/ft2 (107 and 610 g/m2) for different grades and
wire sizes. The galvanizing premium is usually 35 to
40% over non-galvanized wire. Deformed wires are
seldom galvanized.
Where galvanizing of reinforcing bars is desired, galvanizing in accordance with ASTM A767/A767M is
usually performed after fabrication but may be done
before fabrication. The ASTM specification has two
classes of zinc coating weights. Class II (2.0 oz/ft2 [610
g/m2]) normally is specified for precast concrete units.
ASTM A767/A767M requires that all damaged coating
be repaired with a zinc-rich formulation (92 to 95%
metallic zinc in the dry film) in accordance with ASTM
A780/A780M, and sheared ends coated with a zincrich formulation.
Some precasters prefer galvanized reinforcement because it minimizes the potential for staining caused by
red rusting during the prolonged storage of reinforcing
steel, which often is required by the economics of bulk
purchasing.

4

A galvanized coating is anodic to the base steel and
provides cathodic protection to exposed steel, such as
cut edges or holes in the galvanized coating due to
abrasion or impact. Galvanizing is “sacrificial” protection; therefore, in a corrosive environment, it also corrodes. The zinc oxide does not crack the concrete as
quickly as iron oxidation products because it occupies
about one-third less volume for a given weight and is
loose and powdery.
The rate of pressure buildup from corrosion products
is reduced by using a zinc sacrificial barrier, increasing the time before corrosion-related cracking occurs.
Whether cracking of the concrete occurs subsequently
or not depends on many factors such as the strength of
concrete, amount of concrete cover, size of galvanized
member, exposure conditions, and chemical composition of concrete. To what extent the base steel will be
corroded is uncertain because galvanizing does furnish
sacrificial protection to the steel.
Galvanized reinforcement is recommended when
minimum cover requirements cannot be achieved, or
when the concrete is exposed to a particularly severe
environment. However, a detrimental chemical reaction can take place when the concrete is damp and
chloride is present. Therefore, the benefit obtained by
galvanizing is questionable for members subjected to a
marine atmosphere.
EPOXY-COATED REINFORCEMENT. Epoxy-coated
reinforcement has been widely used in aggressive environments since the early 1970s. Epoxy-coated reinforcing bars should conform to ASTM A775 and epoxycoated WWR should conform to ASTM A884. ASTM
A934 covers epoxy coating of fabricated bars. The effectiveness of epoxy coatings for protecting steel from
corrosion has been well documented. However, this
can only be achieved if the coating has minimal pinholes (holidays) and is not significantly damaged. The
corrosion resistance is related to holiday count (that is,
any hole or defect in the coating that permits corrosion
current to pass between the bare steel and liquids) and
extent of damage.
In practice, there will be some pinholes in the coating
when the epoxy-coated reinforcement leaves the factory (ASTM requires not more than an average of one
holiday per foot). Additional damage may occur during
transport to the precast concrete plant, placement of
the reinforcement, and placement and vibration of the
concrete. Such damage should be repaired but locat-

ARCHITECTURAL PRECAST CONCRETE

|

311

4

DESIGN

4.4.7.7 Corrosion Protection / 4.5 Connections

ing all the holidays is difficult and very time consuming. The after-fabrication cut ends and damaged areas
should be patched using the manufacturer’s approved
patch compound.
In addition, bent bars, although visually undamaged
might have lower corrosion resistance than straight
bars. The long-term durability of structures employing
epoxy-coated reinforcement will depend on the progress of corrosion where defects occur in the coating
material. If defects are limited and corrosion does not
spread beneath the coating, long-term performance
should be possible. If high-quality practices cannot
be used, the use of epoxy reinforcement should be
limited.
Epoxy-coated reinforcement does not control cracks
as well as uncoated reinforcement. Cracks in concrete
members reinforced with epoxy-coated reinforcement,
if they occur, will tend to be larger than cracks of similar concrete members reinforced with mild reinforcing
steel. If crack-free design is used, then there is no reason for epoxy coating. Thus, epoxy-coated reinforcement should not be specified for architectural precast
concrete that is expected to be nearly crack free.
Epoxy-coated reinforcement may add significantly to
the cost of the precast concrete products. This results
from epoxy-coated WWR costing two to two-and-ahalf times that of plain WWR, and depends on the size
of the wires. The premium cost of epoxy-coated reinforcing bars ranges from 15 to 30% more per pound
than plain bars.
The use of epoxy-coated reinforcement in architectural
precast concrete panels has not been observed to produce a significantly better performance. This could be
due to the difficulties in achieving the ideal epoxy coating. The reduction of performance, when exposed to
heat or fire, is also a concern. Epoxy coatings usually are
used where the projected loss of bond under high heat
is not significant. Epoxy-coated reinforcement is not
necessary with the environmental conditions typically
experienced by architectural precast concrete as long as
adequate cover and a low w/c concrete are used.

4.5 C ONNECTIONS
4.5.1 G
 eneral
Connections are a significant design consideration
that influence the safety, performance, and economy
of the precast concrete system. Many different con-

312

|

ARCHITECTURAL PRECAST CONCRETE

nection details are often required to accommodate the
multitude of sizes and shapes of precast concrete units
and varying support conditions.
Regardless of whether an architectural precast concrete element is used in a loadbearing or a non-loadbearing application, various forces must be considered
in connection design. In non-loadbearing applications,
a cladding panel must resist its self weight and all other
appropriate forces, such as earthquake, wind, snow,
restraint of volume changes and effects of support system movement, construction loads, loads from adjacent materials, and any other specified loads. These
forces are transferred by the architectural precast concrete element through connections to the supporting
structure. If the panel is loadbearing, then in addition
to the above, each connection must also resist and
transfer dead and live loads imposed on it by floor and
roof elements.
A major advantage of precast concrete construction is
rapid erection. To fully realize this benefit and to maximize economy, field connections should be simple,
repetitious, and easy to install. Precasters and erectors
have developed individual, specific connections over
the years that they favor because they suit their particular production and/or erection techniques.
Connections should comply with local building codes
and also satisfy the functional and aesthetic requirements of the project, such as recessing for flush floors
and/or exposed ceilings. Connections are designed for
each project. Nevertheless, general concepts governing
the design, performance, and material requirements of
connections can be formulated for each project. For
the most effective design, along with efficient connections details, it is recommended that the designer coordinate the connection concepts with a precast concrete manufacturer prior to finalizing the plans.
Terminology: Terms used in this section may be unfamiliar to some. A precast concrete unit (aka panel
or element) is attached to the main building structure (or frame) with a connector or connection that
consists of several parts. The body is the main part
of the connector and may also include fasteners between its components (for example, the angle in Fig.
4.5.24 (page 335) and Fig. 4.5.25 (page 335), the tube
and plate assembly, with the weld being a fastener).
Fasteners are items such as bolts, welds, weld plates,
or even grout or epoxy resin. Fasteners may be used
to attach the body to other portions of the connec-

DESIGN

4.5 Connections / 4.5.2 Design Considerations

tion. The body, if more than very simple, is sometimes
called a bracket or outrigger and according to what
it projects from, a panel bracket or a seat bracket. The
seat (or bearing surface) is that portion of the structure upon which the precast concrete unit’s weight is
supported, and could be a floor or roof slab edge, a
beam, or a seat bracket projecting from a column or
beam. Shear (or weld) plates are welded between
parts of a connection, such as a panel bracket and
seat, primarily to transmit horizontal loads from one to
the other. Tiebacks (push-pull, lateral, stay) are the
connectors that resist forces, primarily perpendicular to
a wall panel, due to wind or out-of-plane seismic and
the couple created by eccentric bearing (or support)
connectors. Tiebacks connect at the structure side to
tieback receiver brackets (or kicker) (for example, the
four angle assembly in Fig. 4.5.33 ( page 336). Anchors
or anchorages are that part of the connection that are
embedded in the concrete, either in the precast concrete unit (panel anchors) or the main structure (structure anchors), and typically are headed studs, bolts,
threaded inserts (coil or national coarse), deformed
bars, or structural shapes. Expansion or chemical anchors are also used at times. Embedments are items,
usually steel fabrications (with anchors), cast into concrete. Adjustable inserts are (usually proprietary) assemblies that have internal adjustability.

4.5.1.1 D
 esign Responsibilities
A successful project requires close cooperation and
coordination between all participants. With the current complexity of construction, it is essential to have
design input by the precaster at an early stage. The
precaster will be able to supply suggestions and designs that ensure that maximum design efficiency is
achieved at the lowest erected cost.
The precaster will usually develop the precast concrete connection details and be responsible for those
details and their performance. Practices regarding the
assignment and acceptance of responsibility in design
and construction vary. It is imperative that the responsibilities of various parties be clearly and firmly established in the contract documents (see Section 4.1 for
further discussion).
The building frame must be designed with adequate
bracing and be stiff enough to support the precast
concrete panels without unanticipated distortion, recognizing that panel loads are usually concentrated at

4

discrete points, which may require localized strengthening and/or stiffening of structural members. The
contract documents should clearly show acceptable locations for bearing connections and structural loading
requirements. When the panel support system requires
projecting brackets, kickers, hardware, and bracing integral to the panel-support frame, they should be designed by the EOR with input from a precaster.

4.5.2 Design Considerations
The primary purpose of a connection is to transfer
loads between the panel and the supporting structure.
In doing so, connections must satisfy design criteria
including:
1. Strength. A connection must have the strength
to safely transfer the forces to which it will be
subjected during its lifetime. In addition to gravity
loads (such as panel self weight), the forces to be
considered may include:
a. Wind, seismic forces, and blast forces.
b. Forces from restraint of material volume change.
c. Forces induced by restrained differential movements between the panel and the supporting
structure.
d. Forces required for stability and equilibrium.
e. Loads from adjacent materials, such as windows,
louvers, signs, and other panels.
2. Ductility. This is the connection’s ability to accommodate deformations without failure. Ductility is
generally achieved by designing connections so
that steel components yield prior to failure of the
concrete.
3. Volume change accommodation. Restraint of
creep, shrinkage, or temperature strains can induce large stresses into precast concrete members
and their supports and thus, must be considered
in connection design. It is usually far better for
the connection to allow some movement to take
place, thus minimizing such stresses. The proper
accommodation of thermal movements in the wall
is a major design consideration. In non-loadbearing units, such movements should be allowed to
take place in individual units with no (or a minimal)
effect on adjacent units. Movements caused by
long-term concrete shrinkage (after a reasonable
curing time) and creep are normally insignificant in
cladding units, but these movements for loadbear-

ARCHITECTURAL PRECAST CONCRETE

|

313

4

DESIGN

4.5.2 Design Considerations

ing panels should be considered in the structural
design. The PCI Design Handbook supplies procedures for estimating such movements.
4. Durability. When exposed to weather or used in
a corrosive atmosphere, connections should be
adequately protected with concrete, paint, galvanizing, or other coatings. Stainless steel is rarely
used for connection hardware because of cost.
All exposed connections should be periodically inspected and maintained.
5. Fire resistance. Connections should be protected
as required.
6. Constructability and economy. The following
items related to constructability and economy should
be considered when designing connections:
a. Use standard connection types.
b. Use standard hardware items and as few different sizes as possible.
c. Avoid reinforcement and hardware congestion.
d. Avoid penetration of forms.
e. Consider clearances and tolerances.
f. Use industry production and erection tolerances.
g. Plan panel connection operations to use a hoist
or crane for the shortest possible time.
h. Provide for appropriate field adjustment.
i. Provide accessibility to complete the connection
from the same floor level.
j. Ensure connections are concealed within space
provided by interior finishes.
k. Minimize length and size of field welds.
Fig. 4.5.1 Forces on a panel subjected to wind suction or
seismic and eccentric loading.
T + T’

T, C = Forces to restrain panels due to
eccentric load
T’ = Force due to wind suction or seismic
W

R = Reaction
W = Weight
- C + T’
R

e

314

|

ARCHITECTURAL PRECAST CONCRETE

7. Wind considerations. Building codes delineate
methods by which design wind pressures should
be determined for any structure. However, these
loads may not be adequate for localized portions
of a tall structure that may be subject to gusting
or funnel effects produced by adjacent structures.
In these cases, a wind tunnel test is normally used
to assess wind loads. The lateral deflection of thin
panels subjected to wind should be determined,
particularly if they are attached to, or include, windows. Wind suction must be considered in design
with the magnitude dependent on the panel shape
or building configuration. Although the design of
the panel itself will generally not be critical for
wind loading, the design of the connections may
be. This is particularly true for tension connections
that resist eccentric gravity loads (Fig. 4.5.1).
8. Seismic considerations. Seismic motions occur
in both directions on all three axes (although the
vertical seismic forces generally do not control) and
the resultant forces may be very large. These forces
are usually reduced by designing structural systems
with sufficient ductility to absorb the seismic energy and dissipate the forces. Seismic effects can
result in substantial story drifts, which are horizontal movements of one floor with respect to those
above and below.
Once the cladding panelization is determined, the
general seismic design approach for architectural
cladding is to determine how the panel will behave
in response to drift and then to configure the connections to accommodate that behavior. Joint locations and widths are important aesthetic concerns
and should be considered in the design process.
For design convenience, all gravity and seismic
forces are assumed to originate at the unit’s center
of gravity. Horizontal forces and movements are
considered one direction at a time; left, right, in,
out, and algebraically added to gravity loads. The
story drift is generally accounted for with connections that flex or slide. For example, a tieback connection is designed to deform under lateral forces,
so it should not transmit racking forces from the
support frame to the panel. All connections need
to be ductile and designed and anchored in such a
manner as to preclude sudden failure.
9. Blast resistance. Blast resistance may be specified
for buildings judged to be a potential target for

DESIGN

4.5.2 Design Considerations / 4.5.2.1 Panel Configuration

attack. Blast loading is characterized by very high
forces acting over very short periods of time—typically measured in milliseconds. Ductile materials and detailing and redundant systems should
be used to prevent the units from becoming a
“fall hazard.” This subject is treated more fully in
Section 5.6 of this manual.

Fig. 4.5.2 Typical arrangement of precast concrete panels.

(a)

(b)

4.5.2.1 P
 anel Configuration
Panel size and the number and spacing of the connection locations all influence connection design. In
general, the minimum number of connections and the
largest size of panel, subject to limits in shipping and
erection, are the most economical.

(c)

The location of joints between precast concrete panels is an important part of the evaluation of economical connections. Figure 4.5.2 shows a few of the many
possible panel configurations for a window wall. In addition to aesthetics, defining the panel configuration
requires recognition of the joint size and movement
with story drift due to wind or earthquake loads. When
possible, it is advantageous to locate panel joints at
gridlines or column lines.
Typical panel connections generally consist of two
bearing connections and four lateral (or tieback) connections. Bearing connections and tieback connections are
sometimes combined. The weight of a panel should be
supported on not more than two points and at one level.
If supported by more than two points or at more than
one floor, the deflections of supporting frame members
may cause the weight distribution to be different than
calculated and could compromise performance.
Figure 4.5.3 illustrates schematically some common
connection locations for different panel types. Figure
4.5.3(a) represents a typical (floor-to-floor) wall unit.
Figures 4.5.3(b) and (c) show possible connection locations for a narrow unit, such as a column cover, and (d)
shows a wide unit, such as a spandrel, with optional,
intermediate tieback connections to minimize flexural
stresses and deflection in longer panels. Spandrel panels should be supported at floor level and restrained at
a column or other vertical member rather than at the
underside of the floor member. This prevents potential
creep rotation of the edge member from affecting the
alignment of the panel. Mid-length tiebacks for spandrels and mid-height connections for taller column
covers are sometimes used to stiffen the panel for lateral loads, resulting in thinner, lighter, and more eco-

4

(d)

(f)

(e)

(g)

Individual Panel

Fig. 4.5.3 Typical cladding panel connector locations.
Column cover options
(b)
(c)

(a) Wall Panel

Bearing Connection
Tieback Connection
(d) Spandrel Panel

Optional
for long panels

ARCHITECTURAL PRECAST CONCRETE

|

315

4

DESIGN

4.5.2.1 Panel Configuration / 4.5.2.2 Panel–connection–structure interaction

nomical panels. Special consideration should be given
to story drifts if connecting units at three levels. In all
cases, the basic connection concepts are similar. In
high seismic areas, force levels often have led to fixing
both bearing connections with no ill effects, in spite of
volume change considerations.
Story drift is less of an issue if spandrel panels’ loadbearing connections and tieback connections are located on the floor beam (Fig. 4.5.3[d]). In this instance
the tiebacks are not affected by story drift because the
top and bottom of the floor beam move together.
Connection details and joint sizes between cladding panels should be designed to accommodate any
shrinkage, story drift, or other expected movement of
the structure, such as sway in tall, slender, steel-frame
structures. Story drift must be considered when determining joint locations and sizes, as well as connection
locations and types. Connections that permit movement in the plane of the panel for story drift (by flexing
of steel or sliding, with slotted or oversized holes, or
other methods allowing for equivalent movement and
ductility) are desirable.
In loadbearing wall construction, horizontal joints
and connections usually occur at, or preferably just
above floor levels and at the foundation or transfer
beams. Joints may be between floors and walls or wall
units only. The principal forces to be transferred are
vertical and horizontal loads from panels above and
from the diaphragm action of floor slabs. When the
panels are “stacked,” one panel bearing on another,
consideration should be given in the connection design to drift.

4.5.2.2 P
 anel–connection–
structure interaction
The interdependency of panel behavior and the supporting frame must be understood.
The way cladding panels behave in response to displacement of the supporting structure can be categorized as shown in Fig. 4.5.4.
In-Plane Translation occurs when the panel is “Fixed”
in-plane to one level. The panel translates laterally with
that level, remaining vertical in elevation. Spandrel
panels and wall panels are typically designed to behave this way.
In-Plane Rotation, also known as “rocking”, occurs

316

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 4.5.4 Modes of panel response to displacement.

Elevation
(a) In-Plane Translation

Elevation
(b) In-Plane Rotation

Section		
(a) Out-Of-Plane Rotation

Section
(b) Out-Of-Plane Translation

when the panel is supported in-plane at two levels of
framing. When the structure displaces, the lateral connections drag the panel laterally, causing it to rotate inplane. This rotation requires bearing connections that
allow lift-off. Column covers are sometimes designed
this way.
Out-of-Plane Rotation is the tilting of a panel perpendicular to its face. This motion is common whenever a
panel is connected to the structure at different levels
of framing. The push-pull connections that support
the panel for wind and seismic forces will also cause
the panel to tilt out-of-plane during story drift. Bearing
connections should be designed to accommodate
out-of-plane roFig. 4.5.5 Corner joint made wider
tation when panto avoid collision.
els behave in this
Seismic
fashion.
Joint

Out-of-plane rotation

Out-of-Plane
Translation is common for spandrel
panels that are attached to a single
level of framing,
such as a perimeter beam and
floor system.

DESIGN

4.5.2.2 Panel–connection–structure interaction

4

collisions occur, possible over-loading of the connections may result as well as cosmetic damage to the
body of the panel. A common way to avoid collisions
is to increase the width of the joint, moving adjacent
panel beyond the limit of movement. A common case
is shown in Fig. 4.5.5 where wall panels form the corner of the building. Wall panels are typically connected
at two framing levels and consequently rotate out-ofplane in response to structure drift, in the case of the
building corner, one panel will rotate while the other
remains vertical resulting in the joint between the two
closing up. To avoid a collision, the corner joint width
must be increased in size in proportion to the anticipated story drift.

Fig. 4.5.6 Joint elevation changes.

Drift

Seismic
joints
(a) Horizontal Joint Transition

Drift

When panels are designed to translate, the horizontal
joint at each level should remain at a constant elevation
whenever possible, as it tracks around the perimeter of
the building. Elevation changes (Fig. 4.5.6) will require
seismic joints at the transitions and detract from the
aesthetics of the cladding.

(b) Preferred Horizontal Joint

When a panel is subjected to an in-plane horizontal
force, its connection system can make the panel rock
up on one corner or translate without tipping or rocking (Fig. 4.5.7 and 4.5.8). It is essential that the poten-

Panel geometry and joints must be configured so
that panels do not collide with one another or with
the supporting structure during a seismic event. When

Fig. 4.5.7 Cladding panel connection concepts - Seismic drift effect (Translating panels).
(a) Wall panels

(b) Spandrel panels

Deflected
position of
grid

Relative lateral
movement

Col. Lines

Seismic force

Seismic
reactions

Floor level

Spandrel panel
(two bearing conn)
C.G.
Seismic reactions

Seismic
force
C.G.

Window

Floor level
Seismic reactions

Note: Gravity and out-of-plane
loads to connectors not shown

Spandrel panel
Bearing Connection
Tie-back Connection
Allowed Movement Direction

ARCHITECTURAL PRECAST CONCRETE

|

317

4

DESIGN

4.5.2.2 Panel–connection–structure interaction

(a)

Grid

(b)

Relative
Lateral
Movement

Floor

Deflected
Position
Seismic
Reaction

Deflected
Position
of Grid

Grid

Fig. 4.5.8 Tall/narrow units.

Floor

Gravity
Reaction

Initial Position
Seismic Load

Floor

Gravity Load

Gravity
Reaction

Bearing
Connection

Seismic
Reactions
Gravity
Load

Gravity Reaction

Floor

Relative
Lateral
Movement

Floor

Tie-Back
Connection

Seismic Reactions
Section A
Section A

Bearing connection
Tie-back connection
Note: All connectors carry out-of-plane loads / gravity and out-of-plane loads to connectors not shown

tial movements be studied and coordinated in regard
to the connection system and the joint locations and
widths. Such considerations may govern the connection design or the wall’s joint locations and widths.
The connection system determines panel movement
in a seismic event. In the case of floor-to-floor wall
panels (Fig. 4.5.7[a]), the upper tiebacks become isolation connections, preventing the building movement
forces from being transmitted to the panel. The panel
is rigidly fixed, to and translates with, the floor beam at
the panel bottom. The seismic force in the plane of the
panel creates a vertical couple and shear forces at the
bearing connections, where these forces and gravity
loads must be resisted. Some designers prefer to provide gravity support for the panel at the top and put
the tieback connections at the bottom, see Fig. 4.5.9.
In Fig. 4.5.8(a) rigid tiebacks at the top of the panel,
together with lift-off allowance at the bottom connections, force the panel to rock when subject to seismic
forces, so its entire weight is being carried on one low-

318

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 4.5.9 Tieback or push-pull connection for a tall precast
spandrel attached to a steel beam. Bearing connection
at top, tieback connection at bottom. Illustration: Chris
Arnold / Tony Alexander, WBDG Building Envelope Study,
National Institute of Building Sciences, Washington, DC.
Bearing Connecton
Precast Concrete
Spandrel Panel

Flexible Connection

DESIGN

4.5.2.2 Panel–connection–structure interaction

er connection. Because the movement occurs in both
directions, each bearing connection must have the capacity to carry the full weight of the element. When
the design intent is to allow rocking of the panels, it is
critical that lift-off be allowed, without tie-down, with
provisions like those shown in Fig. 4.5.67. If lift-off
were prevented, the forces on the bearing connectors
would be increased due to the couple between them.
If the panel is tall and narrow, like a column cover,
bearing connectors can be located so the unit translates with the level of the bearing connectors (Fig.
4.5.3[b] or [c]). If these are vertically close to the panel’s
center of gravity, as in Figs. 4.5.3(c) and 4.5.8(b), the
seismic overturn couple is minimized and the bearing
connectors would carry all gravity and in-plane seismic
loads. The tiebacks would then isolate both the top
and bottom of the panel from their respective floors
(Fig. 4.5.8[b]). An alternative seismic connection sometimes used for tall, narrow units is a single bearing connection, along with sufficient tiebacks for stability.
Historically, in high seismic areas, the most common
application of architectural precast concrete has been
cladding. The International Building Code and the
Uniform Building Code require that wall panels, or similar non-structural elements, be designed to accommodate movements of the structure resulting from lateral
forces and temperature changes. The force requirements
often overshadow the importance of allowing for moisture and thermal volume changes. Panels typically have
two rigid loadbearing connections, with volume change
relief provided only by the ductility of the connections,
and two or more tieback connections with full freedom
of movement in the plane of the panels.
Building codes set the requirements for lateral forces
and story drift accommodation. These parameters,
which are too variable to detail in this publication, depend on seismic design category, building site, occupancy, type, and configuration and should be in the
contract documents. Required drift consideration between floors can be several inches and often present
a greater challenge to the designer than resisting the
forces. This drift requirement is in anticipation of frame
yielding to absorb seismic energy. Connections must
also accommodate the movement during a seismic
event either by sliding or by bending in ductile connectors. Sliding connections must have slots long enough
(after installation tolerance) to account for expected
travel due to story drift without binding or shearing

4

bolts or welds. Flexible connections must have ample
rod or plate length to truly bend and flex under drift
without failing in shear. Careful installation and inspection is required to ensure that tolerances do not negate
the available movement in a way to make the connection ineffective. Weather and corrosion protection
of these sliding connections is also essential to ensure
their long-term performance so the sliding effect can
occur without binding. In the context of seismic connection design, there are four basic connection types:
1. Fixed connections (Fig. 4.5.10a)
2. Rocking connections (Fig. 4.5.10b)
3. Slip connections (Fig. 4.5.10c)
4. In-plane connections (Fig. 4.5.10d)
Connections that resist imposed loads in all directions
are referred to as rigid connections. Rigid bearing connections are generally used in panels that translate inplane as shown in Fig. 4.5.4a.
Bearing connections that allow vertical upward movement (lift-off) are referred to as rocking connections.
This type of connection would allow the panel to rotate in-plane as shown in Fig. 4.5.4b.
Secondary connections that allow horizontal and vertical movement of the panel relative to the supporting structure but resist out-of-plane loading are called
slip connections. Slip connections can be accomplished
by either bolted/slotted arrangements or by flexing of
the connection elements. In-plane connections resist
out-of-plane and in-plane lateral loads, but allows
vertical movement between the panel and supporting
structure.
Connections for loadbearing wall panels are an essential part of the structural support system, and the stability of the structure may depend on them. Loadbearing
wall panels may have horizontal and/or vertical joints
across which forces must be transferred. Loadbearing
panel connections should be designed and detailed
in the same manner as connections for other precast
concrete structural members. It is desirable to design
loadbearing precast concrete structures with connections that allow lateral movement and rotation, and to
design the structure to achieve lateral stability through
the use of floor and roof diaphragms and shearwalls.
Designers are referred to an extensive treatment of design methods in the PCI Manual on Design and Typical
Details of Connections for Precast and Prestressed
Concrete and the PCI Design Handbook.

ARCHITECTURAL PRECAST CONCRETE

|

319

4

DESIGN

4.5.2.3 Tolerances and product interfacing

Fig. 4.5.10 Four basic seismic connection types.

(a) Fixed Bearing

(c) Slip Connection

4.5.2.3 T olerances and product
interfacing
When determining clearance between a new or existing structure and a precast concrete element, dimensional variations of structures—in length, height, and
plumbness—must be considered together with possible deflection, rotation, and irregularities of the supporting surfaces.
Adequate clearance and access must be provided
between precast concrete units and the supporting
structure to allow for product, erection and structure
tolerances. Panel connections often require blockouts
in the edge of floor slabs or access ports in the walls,
which are filled on completion of the connection. In
the case of metal deck systems, size of the blockout
may dictate supplemental deck supports.

320

|

ARCHITECTURAL PRECAST CONCRETE

(b) Rocker Bearing

(d) In-Plane Lateral Connection

It is also important to anticipate geometric variations
in the supporting structure and in the location of hardware in the design of connections (Fig. 4.5.11[a]). In
this case, a preferred clearance of 2 in. (50 mm) is indicated. However, due to allowable tolerances, the actual clearance could vary. The bearing bracket should
be designed for the maximum possible eccentricity.
Connections may be subjected to functional requirements such as recessing for flush floors (Fig. 4.5.11[b])
and/or exposed ceilings. The interior finish systems, furring, and fireproofing must be designed and installed
to conceal the panel connection system. Allowing at
least an extra 1/2 in. (13 mm) between the back of the
drywall and the theoretical back edge of the connection hardware is recommended. Top connections may
be located above the suspended ceiling or covered by a

DESIGN

4.5.2.3 Tolerances and product interfacing / 4.5.2.4 Other detailing information

Fig. 4.5.11 Dimensions for design of connections methods.

2”

(a)
Shim space
1” min
1 1/2” preferred

May be recessed to clear
within interior finish
1 1/2” min clearance
2” preferred
(b)

Note: 1 in. = 25.4 mm

valance. Load support connections at floor level may be
embedded below the floor level or concealed together
with mechanical services. These connections should be
checked early in the process to ensure that recessing
the embed, if necessary, will not interfere with the reinforcement in the structure and that reinforcing details
account for slab recesses.

4

4.5.2.4 Other detailing information
The designer should provide simple and direct load
transfer paths through the connections as well as
ductility within the connections. The number of load
transfer points should be kept to a minimum with
no more than two connections per panel to transfer
gravity loads. Bearing points should occur at the same
level. The EOR should clearly show on the contract
documents the acceptable locations for bearing connections. Details that accommodate adequate production and erection tolerances must be provided as
described in section 4.5.2.3 with the use of slotted or
oversized holes or plates.
Movement allowance in the vertical and horizontal
directions to prevent restraint at tieback connections
can be accomplished by various means. A long, ductile
tieback rod that flexes with drift is common. However,
care must be taken to ensure that this type of tieback
can satisfactorily resist panel forces perpendicular to
the plane of the panel without buckling or excessive
stress. Though costly, low-friction washers or sleeves
slightly longer than the material thickness at a slot
or oversize hole can be used to ensure better performance (Fig. 4.5.66[b]) (page 343). Long or mediumlength rods or bolts may bind instead of slide when
load is applied at the far end. If nuts or bolts are used
at sliding connections, they must be prevented from
either tightening or loosening with movement. This
can be accomplished by using jamb nuts, patent nuts,
punched threads, liquid thread locker, or tack welding
the nut to a square plate washer or a separate stub bar.
If large movement is expected, an articulated tieback
might be considered (Fig. 4.5.66[a]).

All connections should provide maximum adjustability in all directions. Oversized plates, slotted inserts, or
oversized holes in connection hardware can be used to
accommodate plant and field tolerances (Fig. 4.5.14)
(page 329). Welds, shims, and leveling bolts are used
to obtain connection adjustability at the time of installation. The adjustability of the connections facilitates
the alignment of panels and joints.

Twisting and deflection of the panel support beams
can be a concern. Where connections for cladding
transmit panel vertical loads to a steel beam, they
should be centered on the beam unless provisions,
such as bracing, have been made to minimize torsional
stresses in the supporting member. It is usually desirable to locate bearing connections on beams near the
column to minimize beam twisting and deflection.
Where the bearing is on a column, twisting and deflection of the beam is not a problem.

Sufficient space must be provided to make the connections, including room for welding or for turning a
wrench to tighten a nut. When fireproofing is used,
clearances should be planned from the face of the fireproofing material.

Some codes require that the anchorage of the connector be made in such a way as to distribute forces
to the reinforcing steel to avert sudden or localized
failure. The engagement details are left to the precast
concrete engineer. In these situations, some designers

ARCHITECTURAL PRECAST CONCRETE

|

321

4

DESIGN

4.5.2.4 Other detailing information / 4.5.3 Handling and Erection Considerations

use confining hoops, deformed bar anchors, or long
reinforcing bars welded to plates, rather than short
headed studs or inserts in order to better distribute
connection forces to a larger panel area. If anchors are
used near the edge of a precast concrete panel, it is
recommended that they be enclosed in sufficient reinforcing steel, such as hairpins, to distribute any load
back into the panel.
Embedded anchors, inserts, plates, angles, and
other cast-in items should be sufficiently anchored
in the concrete and be detailed so as to not interfere with the mild steel or prestressed reinforcement.
Large reinforcing bar anchors may be impractical due
to their longer required development length, and/or
the difficulty in accomodating the necessary larger
bend radii to interface with the connection hardware.
The design should ensure that interference of reinforcing and connection hardware does not interfere with
concrete placement and consolidation.
In many cases, the wall panels are sufficiently outboard of the supporting frame, to require either outriggers off of the beam (or column) or long panel
brackets (Fig. 4.5.25 through 4.5.28) (page 335). For
seismic forces in the plane of the panel, anchorage of
the longer panel-bearing brackets can become quite
cumbersome when the forces are combined with gravity. If a separate shear transfer plate is added to the
system, such as in Fig. 4.5.49 (page 339), the bracket
anchorage problem becomes more manageable because the shear plate relieves the bracket connection
from carrying some of the seismic forces.
The type of connection shown in Fig. 4.5.11, with
different shapes or angles (gusseted if required), is a
common bearing connection. Such connections are
fastened to the concrete using anchors (Fig. 4.5.65)
(page 342). The connection assembly must be located
accurately in the precast concrete unit to ensure proper
functioning. The anchors should be placed flush with
and perpendicular to the surface.
Bearing pads are sometimes used to distribute loads
over the bearing area and to accommodate construction, fabrication, and erection irregularities. These pads
reduce the concentration of forces at the connection
by deforming readily within their thickness or allowing
slippage. The physical characteristics of bearing pad
materials necessary to satisfy this function are:
1. P ermanence and stability.

322

|

ARCHITECTURAL PRECAST CONCRETE

2. Ability to equalize uneven surfaces and avoid point
pressure.
3. Ability to accommodate movements.
Erection drawings should clearly indicate the type,
location, and orientation of all bearing pads. The pad
supplier or precaster should be consulted when selecting bearing pads. The type of bearing pad required will
depend on the imposed loads and the expected movements of the precast concrete element and support
structure. Two types of bearing pad materials are:
1. Elastomerics with known compression, shear,
and friction properties and ability to deform with
movements.
2. Plastics with low-friction coefficients along with
high compression and shear strength.
If significant movements are expected, soft pads or
low-friction rigid pads should be used. However, if
relative movement is not expected, a bed of rigid material, such as grout or drypack, can be used at the
bearing locations.
Connections cannot be designed without consideration
of the interior finishes and vice-versa. The interior finish
may limit both the type of connection and its location.
In many cases a full-depth blockout (down to a supporting beam) in a floor slab or recessing of a connection
plate is sufficient accommodation for interior finishes.
It is not practical to show the innumerable variations in
connections and finishes that may be encountered.

4.5.3 Handling and Erection
Considerations
In order to achieve the optimum overall economy of
a precast concrete project, it is important to plan for a
minimum of handling of precast concrete components
and erection time at the jobsite. Selection of suitable
connections is also essential.
Allowing the precaster to play an active role in the
development of connection concepts will result in an
optimum solution for achieving efficient erection. It is
important that the designer discuss the use of safe,
efficient, and economical lifting hardware and connections to facilitate erection with the local precast concrete manufacturer and the erector early in the project
planning stage. The precaster or erector may prefer
certain details or procedures not anticipated by the de-

DESIGN

4.5.3 Handling and Erection Considerations

signer. Allowing alternate solutions will usually result in
more economical and better performing connections.
The following is a list of items that should be considered during the selection, design, and detailing of connections to facilitate safe and speedy erection:
1. H
 oisting and setting the precast concrete pieces
is usually the most expensive and time-consuming
element of erection. Connections should be designed so that the erector can safely secure the
member to the structure and release from the
crane in a minimal amount of time. If necessary,
temporary bracing, leveling devices, or connections
can be used, with final adjustment and alignment
in all directions relative to the structure and adjacent components completed after release from the
crane. This is particularly important if the permanent connection requires field welding, grouting,
drypacking, or placement of cast-in-place concrete. Temporary connections should not interfere
with or delay placement of subsequent members.
Temporary connections may have to be relieved or
cut loose prior to completion of the permanent
connections.
2. A
 certain amount of vertical adjustment for alignment at the connection is normal. Precasters have
a number of different methods for alignment. The
common approaches for panel alignment are illustrated in Fig. 4.5.12. The choice between these
procedures and alternates should be the option of
the precaster and/or erector.
Final panel alignment should not be attempted
until all the panels on a beam are in place. As part
of the preliminary layout, the erector should note
the theoretical centerlines of the joints so that the
panels may be centered between these lines, with
minor adjustments to equalize joint widths.
A popular method for vertical adjustment shown
in Fig. 4.5.12(b), uses a leveling bolt. The panel is
usually set a little high and lowered to the correct
elevation with a wrench. The bolts may get very
large for heavy panels and often have a rounded
point to prevent “walking” when turned and to
ease horizontal adjustments. A head-down carriage bolt can also be used.
3. When a unit cannot be erected within the tolerances assumed in the connection design, the
structural adequacy of the connection should be

4

Fig. 4.5.12 Vertical adjustment.
(a)

Shims

(b)
Leveling bolt

checked by the responsible engineer, and if required, the connection design should be modified.
No unit should be left in an unsafe condition. Field
changes to connections that could induce additional stresses in the unit or connection assembly
(other than adjustments within the prescribed tolerances) should only be made after review by the
connection designer.
4. Connections should be planned so that they are
accessible to workers from a stable deck or platform. Erection techniques that require temporary
scaffolding should be avoided, if possible. Room to
turn wrenches should be provided for bolted connections. Foundation piers should extend above
rough grade so that the anchor bolts can be readily
adjusted during erection and so that the base plate
need not be grouted down in a hole, which can
fill with water and debris. Properly drypacking wall
panel bases in a narrow excavation is difficult.

ARCHITECTURAL PRECAST CONCRETE

|

323

4

DESIGN

4.5.3 Handling and Erection Considerations / 4.5.4 Handling and Lifting Devices

5. Connections that serve similar functions within the
building should be standardized as much as possible. Standardization of details facilitates selection
and shipment of connection items to the project.
As workers become familiar with the procedures
required to make the connection, productivity is
enhanced, and the potential for error is reduced.
Whenever possible, items such as field bolts and
loose angles should be of the same size for all connections. Bolts with National Coarse or coil threads
are most commonly used.
6. Connection details should allow erection to proceed independent of ambient temperatures without temporary protective measures. Materials such
as grout, drypack, cast-in-place concrete, and epoxies usually need protection or other special provisions when placed in cold weather. Also, welding is slower and not as simple when the ambient
temperature is very low. If the connections are designed so that these processes must be completed
before erection can continue, the cost of erection
is increased and delays may result.
7. Connections that are not susceptible to damage
in handling should be used. Reinforcing bars, steel
plates, dowels, and bolts (particularly threads) that
project from the precast concrete unit are subject
to damage during handling and should be avoided
whenever possible. Connectors attached to the
panel with threaded inserts may be removed during shipping to minimize damage and facilitate
placement of units on trucks. Threads of inserts
and projecting bolts should be protected from
damage and rust. The precast concrete manufacturer should clean out and cap or seal sleeves or
inserts until used, to prevent dirt and water from
entering and freezing or causing corrosion.
8. When cast-in-place concrete, grout, or drypack
is required to complete a connection, the detail
should provide for self forming, if possible. When
not practical, the connection should allow for easy
placement and removal of formwork. Field patching and finishing should be kept to a minimum.
With steel frames, it may be necessary to determine
how far ahead its final connections and/or floor slabs
must be completed, and what the minimum strength
should be for cast-in-place concrete structures, prior to
loading them with precast concrete units.
In the case of high-rise, cast-in-place concrete struc-

324

|

ARCHITECTURAL PRECAST CONCRETE

tures, delay in erecting precast concrete units will allow
some shrinkage and creep to take place. Connections
and joints should be designed to accommodate shrinkage and creep.
If the structural frame deflection is sensitive to the location or eccentricity of the connection, limits of their
parameters should be given by the EOR.
For loadbearing units, the use of non-metallic shims
or bearing pads, such as neoprene or plastic, may be
used where stresses are not excessive and the risk of
damage to concrete edges is negligible. Any permanent
shims should be non-corrosive or protected so staining
does not occur. Permanent shims should be sized and
positioned so they do not interfere with placement of
joint sealants.

4.5.4 Handling and Lifting Devices
The subject of handling and lifting devices is included
in this section because their design and selection are
determined by requirements similar to those governing
connections.
The design of lifting devices, including checking of
stresses in units during handling, is normally the responsibility of the precaster. Erection inserts cast in
precast concrete members vary depending on the
member’s use in the structure, the size and shape of
the member, and the precast concrete manufacturer’s
preference.
The location of lifting devices can affect the ease of
erection and connection of the precast concrete unit to
the structure. Lifting points should be compatible with
the method of shipping (flat or on edge) and be placed
so that the crane lines do not interfere with the structure.
Lifting devices should be placed on the precast concrete
units so that changes to the rigging can be avoided. The
precast concrete manufacturer should clean and cap or
seal lifting inserts until used by the erector, to prevent
dirt and water from entering and freezing.
The design of lifting devices should consider impact
loads incurred during transit and handling of the units.
The strength of the precast concrete at the time of stripping may only be a fraction of the design strength, and
is a design consideration. Since lifting devices are subject to dynamic loads, ductility should be considered.
Connection hardware should not be used for lifting
or handling unless reviewed and approved by the con-

DESIGN

4.5.4 Handling and Lifting Devices / 4.5.6 Connection Hardware and Materials

nection designer. The final connections should not be
used for erection if they would interfere with the process of attaching the unit to the structure.
Lifting devices can be prestressing strand or aircraft
cable loops that project from the precast concrete, coil
thread inserts, or proprietary lifting hardware. All require engineering and adequate safety factors.
If possible, the placing of lifting and handling devices should be planned so that little or no patching
will be required after use. When temporary lifting and
handling devices are located in finished edges or exposed surfaces, they must be recessed and patched to
match adjacent surfaces. Often, these recesses may be
finished at a later date with prior designer approval.
Specialized lifting equipment may also be used to
eliminate the necessity of patching exposed lifting and
handling devices.
Details of lifting devices should include the consideration of corrosion and possible staining of the finished
product if they are to be left in the units. If the handling
devices interfere with any other function or trade, erection drawings should include removal instructions.

4.5.5 M
 anufacturing Considerations
Because precast concrete plant working conditions,
quality control, and inspection are superior to those in
the field and less dependent on climatic factors, operations demanding high-quality standards are most
efficiently and economically performed in the plant.
Connections should be detailed so that only the least
complicated aspects are completed in the field.
Economy in manufacturing and incorporation of hardware items into precast concrete units demand simplicity and repetition. These demands often result in the use
of one connection detail for several types of units on a
job. Although such connections must be designed for
the most severe load conditions, the cost of the extra
material required will often be less than the detailing,
manufacturing, storing, and scheduling costs for a different connection. As a safeguard against human error,
changes in dimensions and materials used in connection hardware should be in increments large enough
for visual recognition.
Precast concrete units are usually cast in forms that
leave only the top accessible; if one item must be
threaded through and around other items, labor costs
can be significantly increased. When reinforcing bars

4

and connection anchors are concentrated in one location there may be difficulty in placing and consolidating the concrete that can lead to honeycombing unless
special care is taken. Reinforcing bar bend radii, if not
considered, can cause fit problems and leave some regions unreinforced.
A major requirement for the incorporation of hardware in precast concrete units is the positioning of the
hardware to required tolerances. It is often advisable
for the precaster to provide jigs, positioning fixtures, or
mold brackets to ensure proper location, and to maintain tolerances while avoiding skewed or misaligned
hardware.
Hardware placed in precast concrete panels often has
provisions (holes, lugs, and nuts.) so that they may be
secured to jigs, but such details should be left to the
precaster. Hardware projections, which require cutting
through the forms, are difficult and costly to place.
Where possible, projecting hardware should be limited
to the top of the element as cast. Even this placement
increases labor by inhibiting the finishing of the top
surface. The second preference is for the projection to
be on the side forms. The least desirable location for a
projection is on the down or finished face, unless it is
welded or bolted on after stripping from mold.
Avoid casting wood (nailers or lath) in precast concrete. The stress induced by the swelling of the wood
can cause cracking of the concrete.
Dissimilar metals should not be in contact with each
other unless experience has shown that no detrimental galvanic reaction will occur. Casting aluminum into
concrete should be avoided unless a permanent coating dielectrically insulates it, such as bituminous paint or
zinc chromate primer followed by one or two coats of
aluminum metal-and-masonry paint. Also, polyethylene
or similar nonabsorptive tapes or gaskets can be used to
provide local insulation between any dissimilar metals.
Design of connections should consider the ease of
inspection during casting and after completion of installation. Connections should be protected from any
muriatic (dilute hydrochloric) acid used to finish or
clean the units.

4.5.6 Connection Hardware and
Materials
A wide variety of hardware, including reinforcing
bars, deformed bar anchors, headed studs, various in-

ARCHITECTURAL PRECAST CONCRETE

|

325

4

DESIGN

4.5.6 Connection Hardware and Materials / 4.5.7 Corrosion Protection of Connections

serts, structural steel shapes, bolts, threaded rods, and
other materials are used in connections. In order to
achieve strength, the hardware must be properly anchored into the concrete. Plate connections are widely
used in combination with flat metal straps, reinforcing
bars, or metal studs welded to the plate for anchorage.
The exterior surface of the plate is normally flush with
the concrete face and provides a weld area for fastening it to the connector body. By replacing the flat weld
plate with an embedded structural shape, additional
anchorage and strength can be provided. The steel for
anchors should be of a grade and strength similar to
the hardware material in which it anchors, to minimize
welding complications.
Careful placement to required tolerances, including
inclination of protruding shapes, is important because
the bearing surfaces of the panel and the structure
must be parallel in order to obtain optimum bearing
and load transfer. A widely used anchor is the threaded
insert, and numerous variations are available, including
some that are adjustable.

|

4.5.7 Corrosion Protection of
Connections
The need for protection from corrosion will depend
on the actual conditions to which the connections will
be exposed to in service.The corrosion rate of unprotected steel typically ranges from 0.001 in. (0.025 mm)
to 0.005 in. (0.125 mm) per year when exposed to air
and moisture for a significant part of its life.
Connection hardware generally needs protection if
exposed to the weather in service or a corrosive environment. Such hardware should be completely encased in concrete (partially embedded members should
be primed to a depth of 2 in. [50mm]) or otherwise
suitably protected where there is any danger of water
contact. The most common condition requiring protection is exposure to climatic conditions.
Protection may be provided by:

Inserts are easily located and commonly held by bolts
and jigs during casting operations. It is equally important
to place inserts so that the depth of thread is constant
for the same size insert throughout a particular job.
Otherwise, an erection crew may not always engage adequate thread. Also, a typical size and thread depth for
inserts on projects will minimize the possibility of erection
crews using the incorrect size and length of bolts.

1. Paint with shop primer.

Hardware to be placed at the project site should be
detailed with provisions for simple and safe securing
and be oversized to accommodate tolerances in placing. Angles with suitable nail or screw holes for fastening to side forms are superior to flat plates for accurate
placement. Where large horizontal hardware surfaces
are required, it is advisable to provide air release holes
to prevent air pockets and ensure proper concrete consolidation under such surfaces. Hardware for cast-inplace concrete structures should be detailed to provide
the proper anchorage. Anchorage details should allow
placing with reasonable clearance from reinforcing
steel in the structure.

7. Stainless steel.

The cost of field hardware may be reduced if one
item can be used for connecting two adjacent units.
This requires that connections in the precast concrete
units be located close to the edges, which often facilitates production. The practical considerations for
detailing of field hardware are equally applicable to

326

plant hardware and are normally part of the precaster’s
detailing.

ARCHITECTURAL PRECAST CONCRETE

2. Coating with zinc-rich paint (95% pure zinc in
dried film).
3. Chromate plating.
4. Zinc metallizing or plating.
5. Hot dip galvanizing.
6. Epoxy coating.
The cost of protection increases in the order of listing.
Proper cleaning of hardware prior to protective treatment is important. It should be noted that the threaded
parts of bolts, nuts, or plates should be electroplated, not
epoxy coated or galvanized, unless they are subsequently rethreaded prior to use or threads are oversized.
Where connections requiring protection are not readily accessible for the application of zinc-rich paint or
metallizing after erection, they should be metallized
or galvanized prior to erection and the connections
bolted, where possible. If welding is required as part of
the field assembly, the weld slag must be removed and
the weld painted or otherwise repaired to match the
parent material. For galvanized items, the galvanizing
repair paint should be a minimum of 0.004 in. (0.10
mm) thick and conform to ASTM A780.
Special care should be taken when galvanized as-

DESIGN

4.5.7 Corrosion Protection of Connections / 4.5.8 Fasteners in Connections

semblies are used. Many parts of connection components are fabricated using cold-rolled steel or cold
working techniques, such as bending of anchor bars.
Any form of cold working reduces the ductility of
steel. Operations such as punching holes, notching,
shearing, and sharp bending may lead to strain-age
embrittlement of susceptible steels. This is particularly
the case with high-carbon-content steel. The embrittlement may not be evident until after the work has
been galvanized. This occurs because aging is relatively
slow at ambient temperatures but is more rapid at the
elevated temperature of the galvanizing bath.
The recommendations of the American Hot Dip
Galvanizers Association and the practices given in ASTM
A143, Recommended Practice for Safeguarding Against
Embrittlement of Hot Dip Galvanized Structural Steel
Products and Procedure for Detecting Embrittlement,
and CSA Specification G164, Galvanizing of Irregularly
Shaped Articles, should be followed.
Some designers specify the use of stainless steel connections in highly corrosive environments to prevent
long-term corrosion; however, it is extremely costly.
While this may appear to be the best possible corrosion
protection, designers are cautioned that the welding of
stainless steel produces more heat than conventional
welding. The increased heat input, plus a higher coefficient of thermal expansion, may create adverse hardware expansion and greater cracking potential in the
adjacent concrete, thus potentially promoting accelerated long-term deterioration. If welding is to be done
on stainless steel connection plates, edges should be
kept free from adjacent concrete to allow expansion
during welding without spalling the concrete.
Fireproofing of connections may be necessary, depending on codes and/or insurance requirements. In
many cases, fireproofing with concrete cover will also
provide corrosion protection. Many types of connections in precast concrete construction are not vulnerable to the effects of fire and, consequently, require
no special treatment. For example, direct bearing areas between precast concrete panels and footings or
beams that support them do not generally require any
special fire protection, nor do concrete haunches.
If the panels rest on elastomeric pads or other combustible materials, protection of the pads is not generally required because pad deterioration will not cause
collapse. Nevertheless, after a fire, the pads would
probably have to be replaced, so protecting the pads

4

might prevent the need for replacement. If the connections are to be fireproofed or concealed, this fact
should be indicated in the contract documents.
Connections that can be weakened by fire, and thereby jeopardize the structure’s load-carrying capacity,
should be protected to the same degree as required
for the structural frame. If, for example, an exposed
steel bracket supports a precast concrete element that
is required to have a designated fire rating, the steel
bracket must have the same fire rating.
Many connections simply provide stability and are under little or no stress in service. While fire could substantially reduce the strength of such a connections, no fire
protection is necessary. Connections that have steel elements encased in concrete, drypacking, or grout after
erection usually need no additional protection.
There is evidence that exposed steel hardware used in
connections is less susceptible to fire-related strength
reduction than other steel members. This is because
the concrete provides a “heat sink,” which draws off
the heat and reduces the temperature of the steel.
Fireproofing of connections is usually accomplished
with sprayed cementitious fireproofing, or mineral fiber, intumescent mastic compounds, or enclosed with
gypsum wall board.
Figure 4.5.13(a) shows the thicknesses of various,
commonly used, fire-protection materials required for
fire endurances up to four hours when applied to connections comprised of structural steel shapes. The values shown are based on a critical steel temperature
of 1000 °F (538 °C) (that is, a stress-strength ratio of
about 65%). The values in Fig. 4.5.13(b) are applicable
to concrete or drypack mortar encasement of structural steel shapes used as brackets.
When a rational analysis or design for fireproofing is
not performed and concrete is used to fireproof the
connections in the field, a conservative estimate would
suggest that such concrete should have a thickness in
inches corresponding to the specified hours of fire rating. Unless the nature of the detail itself supports such
concrete, it should be reinforced with a light weldedwire reinforcement.

4.5.8 Fasteners in Connections
Shim stacks are used to transmit bearing loads between elements or to the supporting structure. Shims
can be made from plastic, steel, neoprene, or other

ARCHITECTURAL PRECAST CONCRETE

|

327

4

DESIGN

4.5.8 Fasteners in Connections

Fig. 4.5.13 Thickness of fire protection materials on steel connection components.
4
b ≥ 12"

b = 8"

t = 5/16"

F
SM

SMF
or V
C

M

VCM

3

t = 5/16"
2

b

1

0

Concrete or Dry-Pack Mortar
b=M
 inimum Width of
Concrete Protection

t = Minimum Thickness for Steel
Subjected to Fire from Both Sides

IM

Fire Endurance, Hr

t = 11/16"

1

2

3

4

0

1

2

3

4

Thickness, Protection Material, In.
(a)

(b)

(IM = Intumescent Mastic, SMF = Sprayed Mineral Fiber, VCM = Vermiculite Cementitious Material)

suitable materials and are often used as spacers or as
means of leveling or aligning adjacent components.
Shim stacks often lead to multiple shims, which are
not suitable for transmitting lateral forces. Shims range
in thickness from 1/16 in. to 2 in. (1.6 to 50 mm). Onepiece shim blocks may be used and are favored by
some erectors when they are combined with reliable
panel dimensions and a survey of the structure prior
to the placing of panels. When shim stacks are to carry
loads permanently it may be advisable to weld them
together, especially if they are used at a sliding connection. Such stacks might need to be supplemented
with separate shear plates similar to those in Figs.
4.5.45 through 4.5.51 (pages 338 to 339) for horizontal loads. Temporary shims should be removed from
joints of non-loadbearing units after connections are
completed and before applying sealant.

328

|

Note: 1 in. = 25.4 mm

release of the unit from hoisting equipment should be
minimized. If structural considerations indicate the necessity for welding, the precast concrete erector may
temporarily support and brace the unit so the crane
will not be tied up. Bolted connections minimize this
concern. Provisions must be made to maintain the unit
safely in place until final connections are completed.
Welding should be performed in accordance with the
erection drawings by personnel that have been certified
for the welding procedures specified. These drawings
should clearly specify type, size, length and location of
welds, and any critical sequences. All welding, including tack welds, should be made in accordance with the
applicable provisions of the American Welding Society.
It is both difficult and expensive to weld overhead or in
confined places, so these situations should be avoided.
Welding on galvanized hardware requires proper procedures to avoid contamination due to poor-weld quality.
Cold galvanizing, zinc-rich paint should be applied over
welded areas to replace the removed galvanizing.

Welded connections for cladding panels are structurally efficient and easily adapted to varying field
conditions. Their strength depends on reliable workmanship and the compatibility of welding materials
with the metal to be joined. Welded connections can
be completed only after final alignment. Where only
a few field connections are to be welded, it is usually
more economical to use an alternate connection rather
than require another trade (welders) on the project.

When welding is performed on components that are
embedded in concrete, thermal expansion and distortion of the steel may destroy the bond between the
steel and concrete or induce cracking or spalling in the
surrounding concrete.

Hoisting and setting time is critical for economical
erection. Welding that must be executed prior to the

The extent of cracking and distortion of the metal is
dependent on the amount of heat generated during

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.5.8 Fasteners in Connections

welding and the stiffness of the steel member. Heat
may be reduced by:
1. U
 se of low-heat welding rods of small diameter.
2. U
 se of intermittent, rather than continuous, welds.
3. U
 se of smaller welds and multiple passes.

4

Fig. 4.5.14 Slotted or oversize holes.
(a) Two-axis adjustability
during erection. Welding
square plate washer after
alignment limits movement to one axis.

Using thicker steel sections can minimize distortion. A
minimum of 1/4 in. (6 mm) is recommended for plates.
Some precasters line the metal with sealing foams,
clay, or other materials to minimize the risk of the concrete cracking, especially if the metal is thinner than 3/8
in. (10 mm).
Cracks in concrete may be wider during welding, but
will usually close significantly after the steel cools.
Bolted connections often simplify and speed up the
erection operation because the connection is positive
immediately. Final alignment and adjustment of the
panel can be made later without using valuable crane
time. In using bolted connections, it is desirable to standardize the size of attachment hardware (clip angles
and bolts). Standardization minimizes errors and inventory of hardware and improves productivity. With bolted
connections, 3/4, 1, or 11/4 in. (19, 25, or 32 mm) diameter bolts with National Coarse or coil threads are considered standard in the precast concrete industry and
should be used when possible. Coil thread stock or coil
bolts are often used in lieu of National Coarse threaded
stock or bolts to minimize the time required to make
connections and reduce the risk of thread damage.
Bolted connections should allow for industry erection
tolerances. Slotted or oversized holes should be provided to accomodate variation and tolerance when they
do not conflict with the design intent, Fig. 4.5.14.
Precautions must be taken to ensure adequate engagement of threads in all threaded inserts. Following
erection of a precast concrete unit where slotted connections are used, a check of bolt position and tightness should be made. For sliding connections, the bolt
should be properly secured, with lock washers, tack
welding, or other means to prevent tightening or loosening. They should be snug, but not so tight that they
cannot move within the slot. Low friction washers
(Teflon® or nylon) may be used to ensure movement
capability. Roughness at sheared or flame-cut edges
should be removed. Washers, when used, should be
large enough to overlap the sides of slots or oversize
holes, and allow for full movement. Plate washers with
off-center holes allow maximum flexibility without re-

(b) Alternate adjustable insert
Two axis adjustability via
slotted insert and slotted
washer with perpendicular
axes

Obstruction
(c) Off-center hole in square
plate washer takes full advantage of oversize hole,
yet allows maximum weld
length in spite of varying
dimension to an obstruction during alignment

2”

1”

3”

4”

quiring separate size parts (Fig. 4.5.14[c]). Adjustable
inserts can also allow for movement.
When the connection cannot be made because the
insert is out of place or missing, the connection designer should approve any modifications.
Expansion anchors are often used as connections
at foundations or for corrective measures when castin inserts are mislocated or omitted. They are inserted
into drilled holes in hardened concrete. Performance
of these anchors is dependent on the quality of field
workmanship. Strength is obtained by tightening of
the bolt or nut, thus expanding parts of the anchor,
which exert lateral pressure on the concrete. Types that
increase, rather than reduce, expansion under load are
preferable. Anchorage strength depends entirely on
the expansion force. For connection reliability, the importance of correct installation and quality control cannot be overemphasized. Holes for expansion anchors
must be drilled straight, deep enough, with the proper
diameter, and be cleaned out. The minimum distance
to the edge of the concrete and spacing between

ARCHITECTURAL PRECAST CONCRETE

|

329

4

DESIGN

4.5.8 Fasteners in Connections / 4.5.9 Supply of Hardware for Connections

bolts should be based on the anchor manufacturer’s
recommendations.
The performance of expansion anchors when subjected to stress reversals, vibrations, or earthquake loading
is such that the designer should carefully consider their
use for these load conditions. These anchors should
meet the requirements of ACI 318, Appendix D.
Chemical anchors (resin capsule or epoxy anchors)
could be considered for corrective measures or for
heavy loads. However, chemical anchors may degrade
at temperatures in the 140 to 150 °F (60 to 66 °C)
range. Such temperatures may be experienced in
warm climates, particularly in façade panels with dark
aggregates. Chemical anchors may not be allowed in
fire-rated connection assemblies. Manufacturers’ recommendations and local code provisions for installation must be followed.
Grouting or drypacking of connections is not widely used, apart from base plates or loadbearing units.
The difficulty in maintaining exact elevations and the
inability to allow movements and still maintain weather tightness must also be considered. Grouting should
be used carefully when installed during temperatures
below or near freezing. Units with joints that are to be
drypacked are usually supported with shims or leveling
bolts until drypack has achieved adequate strength.
Shims used for this purpose could be subsequently
removed to prevent them from permanently carrying
the load or to facilitate joint sealant installation. A drypacked joint requires a joint wider than 1 in. (25 mm)
for best results.
Grouted dowel/anchor bolt connections depend
on their diameter, embedded length, and bond developed. Placement of grout during erection usually
slows down the erection process. Any necessary adjustment that is made after initial set of the grout may
destroy bond and reduce strength. It may be better to
provide supplemental bolted connections to expedite
erection.
Erection drawings should show the required grout
strength:
1. B
 efore erection can continue.
2. B
 efore bracing can be removed.
3. A
 t 28 days.
Pressure grouting has been successfully employed in
the joints in a number of cases. In order to contain the pres-

330

|

ARCHITECTURAL PRECAST CONCRETE

sure grout, special neoprene gaskets with suitable vents
are usually designed specifically for the project. Grout is
usually injected from the bottom to minimize voids.
Post-Tensioning may be used to make field connections between precast concrete members using either
bonded or unbonded tendons installed in preformed
voids or ducts. Bonded tendons are made monolithic
with the member and protected from corrosion by
grouting after the stressing operation is completed.
Unbonded tendons are protected against corrosion by
a properly applied preventive coating. The unbonded
tendons are connected to the member only through
the anchorage hardware, which also must be protected from corrosion and fire where appropriate. In the
right circumstances, unbonded tendons can absorb
substantial seismic energy.
Erection drawings should show post-tensioning details and instructions. Bracing may need to remain until
tensioning is complete.

4.5.9 Supply of Hardware for
Connections
The responsibility for supplying field hardware to be
placed on or in the structure in order to receive the
precast concrete units depends on the type of structure and varies with local practice. The furnishing and
placing of the hardware should be clearly defined in
the bid and contract documents. Hardware should be
incorporated into the structure within specified PCI
tolerances according to a predetermined and agreed
upon schedule to avoid delays or interference with the
precast concrete erection and the project schedule.
1. Building frame of structural steel. Projecting
brackets, kickers, hardware, and bracing for precast concrete connections that are integral to the
frame are preferably supplied and installed as part
of the structural steel contract for reasons of economy (minimizes the cost of crew and equipment
standby). This requires sufficient coordination time
to provide proper hardware locations before detailing and fabrication. It is recommended that
the precaster be brought “on board” before the
structural steel is bid, to allow for the design and
detailing of the precast concrete connections to be
incorporated into the steel bid. If added later, these
items are likely to cost more than if included in the
original steel bid.

DESIGN

4.5.9 Supply of Hardware for Connections / 4.5.10 Connection Details

2. Building frame of cast-in-place concrete. Field
hardware may be supplied by the precaster, the
GC, or be a part of the miscellaneous steel subcontract. It is recommended that the precaster
supply this hardware, because this method usually
reduces the problems of detailing and coordination. Hardware is placed by the GC, or the concrete subcontractor, to a layout drawing prepared
by the precast concrete manufacturer. In any case,
award of the precast concrete contract must be
timely to allow design, detailing, and fabrication
of embedded items to be in sequence with casting
schedules.
If anchors or other structural items shown on the
precast concrete manufacturer’s approved drawings cannot be accommodated because of mislocation, unforeseen reinforcing complications, or
the work of other trades, the GC shall obtain approval for any modifications by both the EOR and
the precast concrete manufacturer to ensure the
required structural adequacy of the hardware.
Field-installed, pre-erection hardware consists of
miscellaneous loose steel pre-welded or pre-bolted to
the structure prior to beginning the panel installation.
It is customary for the precast concrete manufacturer
to supply this erection hardware even when not performing the actual erection.

There are many possible combinations of anchors,
plates, steel shapes, and bolts to form various connection assemblies. The details and final assemblies
selected should be optimized considering design criteria, production and erection methods, tolerances,
and economy. Common practice by precast concrete
manufacturers in a given area may also influence the
final selection of details on a particular project. The
connection details are not numbered in any order of
preference.
It is not the intent to limit the type of anchorage of any
connector to the precast concrete to that shown in the
figures. A variety of anchors are shown in Fig. 4.5.65,
which are generally interchangeable and must be integrated with the reinforcement. This is an engineering
task required for each individual project. The details
may sometimes have to be combined to accomplish
the intended purposes. For example, Fig. 4.5.15 and
Fig. 4.5.17 are often combined, and Fig. 4.5.46 shows
how connector anchor loads can be minimized.
All connections must consider tolerances as outlined
in Section 4.5.2.3.
The examples shown cover the following broad
categories:
Fig. 4.5.15 to 22

Direct bearing

DB 1-8

Fig. 4.5.23 to 28

Eccentric bearing

EB 1-6

Erection hardware is the loose hardware needed in
the field for final connection of the panel and is normally supplied by the precaster.

Fig. 4.5.29 to 36

Welded tieback

WTB 1-8

Fig. 4.5.37 to 44

Bolted tieback

BTB 1-8

Fig. 4.5.45 to 51

Shear plate

Accessory hardware, if required to be cast into the
precast concrete units (such as electrical boxes, conduit, window inserts, fastenings for other trades, and
dovetails for flashing), should be designed and supplied by the trade requiring them. If inserts are used,
their locations should be given on the approved shop
drawings of both trades.

Fig. 4.5.52 to 55

Panel to panel alignment

PPA 1-4

Fig. 4.5.56 to 61

Column cover

CC 1-6

Fig. 4.5.62

Beam cover

BC 1

Fig. 4.5.63

Soffit hanger

SH 1

Fig. 4.5.64 to 69

Special conditions

Fig. 4.5.70 to 74

Bearing wall to foundation BWF 1-5

4.5.10 C onnection Details

Fig. 4.5.75 to 77

Slab to bearing wall

Fig. 4.5.78

Slab to side wall

SSW 1

Fig. 4.5.79

Wall to wall

WW 1

This section shows typical details for some of the more
commonly used connections for cladding panels and
loadbearing precast concrete walls, as well as other connections that may be useful in special applications. The
details included are not exhaustive. They should not be
considered as “standard,” but rather, as concepts on
which to build. Detailed design information, such as
component sizes, weld and anchorage lengths, joint sizes, and bearing pad thicknesses is purposely omitted.

4

SP 1-7

SC 1-7
SBW 1-3

Bearing (direct and eccentric) connections are
intended to transfer vertical loads to the supporting
structure or foundation. Bearing should be provided
at no more than two points per panel, and at just one
level of the structure. Bearing can be either directly in
the plane of the panel along the bottom edge, or eccentric using continuous or localized reinforced con-

ARCHITECTURAL PRECAST CONCRETE

|

331

4

DESIGN

4.5.10 Connection Details

crete corbels or haunches, cast-in steel shapes, or attached panel brackets. Transfer of forces perpendicular
to the panel is provided by various tieback arrangements. Adjustability in the support system generally
necessitates the use of shims, leveling bolts, bearing
pads, and oversized or slotted holes.
Direct bearing connections are used primarily for
panels resting on foundations or rigid supports where
movements are negligible. This includes cases where
panels are stacked and self supporting for vertical loads
with tieback connections to the structural frame, floor,
or roof to resist forces perpendicular to the panel.
Eccentric bearing connections are usually used for cladding panels when movements of the support system are
possible. Cladding panels are, by definition, fastened
to and/or supported by a structure located in a different plane. Eccentric bearing connectors (corbel or panel
bracket) cause permanent bending stresses in the supported panel that must be accommodated. Concrete
haunches or corbels also provide a solution for heavy
bending within the panel. Bending combined with tension, shear, and torsion may have to be resisted by the
connection and, in turn, the structure, depending on
the type of connection and load transfer details.
If leveling bolts and shear plates are used, the shear
plates are proportioned for all lateral loads (Fig. 4.5.25).
The leveling bolt is usually left in place to carry the vertical load. If shims are used instead of leveling bolts, and lateral loads are to be carried, a weld plate is recommended,
because the welding of shim edges is usually unreliable
for transmitting significant forces. The erector’s individual
preference for shims or leveling bolts should be allowed.
Bearing connections are usually, but not always, combined with tiebacks.
Tieback (welded or bolted) connections are primarily intended to keep the precast concrete unit in a
plumb position and to resist wind and seismic loads perpendicular to the panel. Welded tiebacks often require
temporary bracing during alignment. Tiebacks may be
designed to take forces in the plane of the panel, or
isolate them to allow frame distortions independent
of the panel and allow movement vertically and/or
horizontally.
Shear plates are generally welded and serve primarily to provide restraint for longitudinal forces in the
plane of the panel. They usually also carry loads perpendicular to the panel, acting as a tieback connection

332

|

ARCHITECTURAL PRECAST CONCRETE

as well. Because seismic force is the most common inplane force, these plates are sometimes referred to as
seismic shear plates. It is, in many cases, uneconomical
to carry longitudinal forces on longer panel brackets of
eccentric bearing connections because their anchorage
loads become very high. In such cases, the shear plate
connection is used to reduce the load on the anchors
(Fig 4.5.46). Longitudinal force transfer on spandrels,
for example, can be accomplished near mid-length of
the member to minimize volume change restraint forces
that would otherwise be additive to the longitudinal
seismic forces.
Panel-to-panel alignment connections are used
to adjust precast concrete units’ relative positions with
respect to adjacent units; they do not usually transfer
design loads. Out-of-plane alignment of panels is sometimes necessary, especially if they are very slender and
flexible and have warps or bows prior to erection.
Column and beam cover connections are used
when precast concrete panels serve as covers over steel
or cast-in-place concrete columns and beams. The cover
units are generally supported by the structural column
or beam and carry no load other than their own weight,
wind, and seismic forces. The weight of a column cover
section is normally supported at one level. Tieback connections for lateral load transfer and stability occur at
multiple levels. Connections must have sufficient adjustability to compensate for tolerances of the structural
system. Column cover connections are often difficult to
reach, and once made, difficult to adjust. For thin flat
units, when access is available, consideration should be
given to providing an intermediate connection for lateral support and restraint of bowing. “Blind” connections, made by welding into joints between the precast
concrete elements, are sometimes necessary to complete the final enclosure.
Soffit hanger connections can be made by modifying many of the tieback connections previously discussed. If long, flexible hanger elements are used, a lateral brace may be provided for horizontal stability.
Special conditions are presented in Figs. 4.5.64
through 4.5.69. These are suggested to help solve
unique or difficult situations.
Bearing wall connections are divided into categories:
those that support the bearing wall and floor or roof
slabs, and those with (non-supported) edges of floor or
roof slab running alongside them. These conditions are
not the same as the connection of an architectural panel

DESIGN

4.5.10 Connection Details

to the structure like the others in this section. They are
included because they often occur in loadbearing wall
panel systems. Many of the tieback, shear plate, and
panel-to-panel alignment connections in Figs. 4.5.29 to
4.5.45 could be used in bearing walls.
Bearing wall to foundation connections and the
direct bearing connections in Figs. 4.5.15 to 4.5.19 are
primarily intended to transfer their gravity loads to the
panel below or to the foundation, although they can usually carry lateral loads, as well. The connections should
provide a means of leveling and aligning the panel. The
attachment method should be capable of accepting the
base shear in any direction. In cases where an interior
core carries lateral loads, this may be accomplished with
a simple welded connection.
Slab to bearing wall connections are used to join
precast or cast-in-place concrete floor or roof members
to precast concrete bearing walls. They transfer any vertical load from the horizontal system and, sometimes,
diaphragm action and on rare occasions provide moment resistance.
Blockouts in wall panels or spandrels as in Figs. 4.5.75(e),
4.5.75(f), and 4.5.76 decrease eccentricity and bending
in the wall panel. Using blockouts in a spandrel would
reduce the torsion stresses and twist during erection. If
discontinuous pockets are used, they require substantial
draft on their sides (1/2 in. [13 mm] every 6 in. [150 mm]
depth to allow blockout stripping) and should have at
least 21/2 in. (63 mm) cover to the exposed face. More
cover (3 in. [75 mm] minimum) is required if the exterior
surface has an architectural finish. In the case of a fine
textured finish, there may be a light appearing area (the
approximate size of the blockout) that shows on the face
of the panel due to differential drying. This may be quite
noticeable, despite the uniformity of the finish. The initial cure of the 2 1/2 to 3 in. (63 to 75 mm) of concrete
versus 8 to 9 in. (200 to 225 mm) in the surrounding
area will make the difference.
When the slab functions as a diaphragm, the connections must transmit diaphragm shear and chord forces
to a structural core, thus reducing the load on individual
exterior walls or spandrel units and their connections.
When the slab-to-wall connection is accomplished with
composite topping, temporary connections or bracing
may be necessary during erection.

4

by a judicious use of bearing pads or weld plates.
Slab-to-side wall connections along the (non-bearing) sides of floor or roof slabs may be required to transmit lateral (diaphragm) loads and should either allow
some vertical movement to accommodate camber and
deflection changes in the floor units, or be designed to
develop forces induced by restraining the units.
Wall-to-wall connections are primarily intended to
position and secure the walls, although with proper design and construction, they are capable of carrying lateral loads from shear walls or frame action as well. The
two locations of wall-to-wall connections are horizontal
joints (usually in combination with floor construction)
and vertical joints.
The most practical connection is one that allows realistic tolerances and ensures immediate transfer of load
between panels.

Fig. 4.5.15 Direct bearing (DB1).
• Lateral restraint not provided
• Has large tolerance
• Finish joint with drypack or sealant
Shim stack

Fig. 4.5.16 Direct bearing (DB2).
• Insert must be jigged plumb
• Allows vertical adjustment without crane
• Finish joint with drypack or sealant
• Bolt head may be welded for tensile or
shear capacity
• Plate may be eliminated, but adjustment
becomes more difficult
• May be inverted with insert below.

Most designs result in some degree of fixity for these
connections. However, a fully fixed connection is generally not desirable. The degree of fixity can be controlled

ARCHITECTURAL PRECAST CONCRETE

|

333

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.17 Direct bearing (DB3).

Fig. 4.5.20 Direct bearing (DB6).

• Reasonable tolerance

• For shaped panels: can eliminate dead load overturn if
shims in line with panel center of gravity

• Provides lateral restraint
• Realignment not possible after
connection complete

• Complex forming, especially if location of haunch changes
• Forming simplified if a bolt-on steel haunch is used

• Requires shims until grouting or
drypacking is done

Haunch

• Cold weather may be a problem
with grouting or drypacking

Shim stack

• Grout could be injected through
tubes, allowing more time for
alignment
• Void may be formed or field drilled
• Finish joint with drypack or sealant

Fig. 4.5.18 Direct bearing (DB4).

Fig. 4.5.21 Direct bearing (DB7).

• Reasonable tolerance

• Preferable if column bearing bracket
shown on contract drawings and
shop-installed

• Provides lateral restraint
• Realignment not possible after connection
complete
• Requires shims until grouting or drypacking
is done
• Cold weather may be a problem with
grouting or drypacking
• Grout could be injected through tubes,
allowing more time for alignment

• Cost substantially more if
bracket field-installed, which
also requires field layout
• Leveling bolt could be used in
lieu of shims
• Can be used in pocket farther
up panel away from joint

• Upper void difficult to fill
• Upper void could be continuous or
intermittent
• Finish joint with drypack or sealant

Fig. 4.5.19 Direct bearing (DB5).

Fig. 4.5.22 Direct bearing (DB8).

• Full strength of bar can be achieved
with proprietary grouted sleeve

• Lateral restraint could be provided
by welding bolt head to seat

• Small tolerance requires jigging

• Could use threaded insert in
lieu of angle assembly

• Requires shims until grouting or
drypacking is done
• Joint may be drypacked or grouted at
same time as sleeve
• Smooth or corrugated sleeve could
replace proprietary sleeve for lower
capacity
• Finish joint with drypack or sealant
• Sleeve can be in either panel

334

|

ARCHITECTURAL PRECAST CONCRETE

Anchor beyond

DESIGN

4.5.10 Connection Details

Fig. 4.5.23 Eccentric bearing (EB1).

Fig. 4.5.26 Eccentric bearing (EB4).

• Coordinate with GC for placement of seat

• Coordinate with GC for placement of seat

• Could use leveling bolt or shims

• Any structural shape could be used for
projecting bracket—if unsymmetrical,
consider torsion

• Could use thicker angle and delete gusset

Make shim stack

• Could eliminate projection from panel by
attaching angle with inserts or welding to
flush plate

• Many types of panel bracket anchorage
could be used

Fig. 4.5.24 Eccentric bearing (EB2).

Fig. 4.5.27 Eccentric bearing (EB5).

• Coordinate with GC for placement of seat

• Same panel bracket can be used with any column size

• Complex haunch reinforcement

• Any structural shape could be used for projecting bracket

• Complex forming, especially if
location of haunch changes

Oversize hole or
vertical slot
in angle

• Haunch could be cast first and
Plate
set in form
washer
• Haunch could be intermittent
or continuous

4

• Many types of panel
bracket anchorage
could be used
Any of the members
shown could be
other structural
shapes.

• Plate washer may require
welding for lateral loads

On both sides of column

Fig. 4.5.25 Eccentric bearing (EB3).

Fig. 4.5.28 Eccentric bearing (EB6).

• Keep bearing at center of beam to avoid torsion

• Same panel bracket can be used with any column size

• Leveling bolt saves time

• Thin tube may require reinforcing plate at bearing

• Could use shims in lieu of leveling bolt
• May require blockout in
floor slab

Coupling
nut

• Different tieback could
be used in lieu of shear plate

Shown with optional shear plate

ARCHITECTURAL PRECAST CONCRETE

|

335

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.29 Welded tieback (WTB1).

Fig. 4.5.32 Welded tieback (WTB4).

• Consider beam deflection

• Consider deflection of support

• Stagger anchor studs to minimize
magnification of force on them
due to variation of shear plate
location

• Slots and bolts allow fast erection—weld after alignment

• Requires bracing until welded
• May also serve as shear plate

Fig. 4.5.30 Welded tieback (WTB2).

Fig. 4.5.33 Welded tieback (WTB5).

• Requires bracing until welded

• Consider deflection of support

• Alignment and welding must
be done before upper
panel is erected

• Slots and bolts allow fast erection—weld after alignment

• Difficult to inspect
• May also serve as
shear plate

Oversized Hole

Fig. 4.5.31 Welded tieback (WTB3).
• Buckling of rod must be considered if
compression load is expected
• Requires bracing until welded

Fig. 4.5.34 Welded tieback (WTB6).
• Coordinate with GC for placement of insert
Plain Rod
with Thread
at One End

• Do not over-tighten threaded rod if
movement in slotted insert to be
allowed
• Slotted bar may be used to fit
proprietary slotted embedment

336

|

ARCHITECTURAL PRECAST CONCRETE

• Adjustment limited by thread length of insert and bolt
• Need adequate clearance for welding
• Weld not required for
compression only
• Could reverse with
plate in structure,
and insert in panel

DESIGN

4.5.10 Connection Details

Fig. 4.5.35 Welded tieback (WTB7).

Fig. 4.5.39 Bolted tieback (BTB2).

• Anchorage of plate and angle could vary

• Alignment can be completed after release from crane

• Shear plate configuration to be determined by load type

• Slots in embedment and angle to be perpendicular to each other
for three-way adjustment

4

• Threaded insert can be used if angle has oversize
hole and plate
washers

Fig. 4.5.36 Welded tieback(WTB8).

Fig. 4.5.39 Bolted tieback (BTB3).

• Oversize hole in angle

• Horseshoe shims allow adjustment perpendicular to panel

• Plate washer could be welded and slotted to control directional
movement See Fig. 4.5.14 for reference

• Oversize hole and plate washer allows adjustment parallel
to panel
• Do not over-tighten bolt if
movement to be allowed
• Plate washer could be
welded and slotted to
control directional
movement.
See Fig. 4.5.14

Fig. 4.5.37 Bolted tieback (BTB1).

Fig. 4.5.40 Bolted tieback (BTB4).

• High-strength rod is advantageous

• Coordinate with GC for placement if insert is used

• Rod flexes for in-plane
movement

• Edge distance and reinforcing in
floor/foundation must be considered

• Bucking of rod must
be considered if
compression load
is expected

• Angle has slotted holes

• Oversize hole
primarily for
tolerance

Insert

ARCHITECTURAL PRECAST CONCRETE

|

337

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.41 Bolted tieback (BTB5).

Fig. 4.5.45 Shear plate (SP1).

• Basically an alternate to BTB1 where long rod
cannot be accommodated

• Primarily for in-plane lateral force
• Also takes out-of-plane force

• Oversize hole both for tolerance
and movement allowance

• Normally one used near center of
panel, with larger panel to beam dimension, so force needn’t be restricted by
long panel brackets

• Tieback rod receiver could be
many configurations

• Trapezoidal plate may be assumed fixed
at beam and pinned at panel to minimize
panel plate anchorage
• Installed after panel fully aligned, so
temporary tieback may be required

Fig. 4.5.42 Bolted tieback (BTB6).

• Thin plate allows some vertical
movement

• Sleeve in concrete column or wall
must be large enough for adjustment
• Bearing pad need not be
adjacent to tieback rod

Fig. 4.5.46 Shear plate (SP2).

• Special care required to
maintain tolerance

• Similar to SP1 except
combined on bearing
connector anchor plate

Bearing pad

• Eliminates need for shear
plate on bearing bracket
• Panel plate anchorage
requirement is lower than if
in-plane force were resisted
by bracket

Fig. 4.5.43 Bolted tieback (BTB7).
• May require bracing until floor is cast

Fig. 4.5.47 Shear plate (SP3).
• Convenient at mid-height
of column covers
• Can be used for rocking or
translating unit, depending
on balance of connection system
• Use in pairs or weld to column

Fig. 4.5.44 Bolted tieback (BTB8).
• Blind connection
• Panel face does not need patching

Cast-in-place or
masonry wall

• Large opening required for access

Temporary Tie

Fig. 4.5.48 Shear plate (SP4).

• If angle is field welded, smaller
access hole allowed, but
temporary bracing required

• Shims carry full weight of panel
• Shims should be adjacent to shear
plate (angle)

• Field tolerances critical

• Angle orientation gives high
capacity in all three axes
• Cannot be installed until after
alignment, so temporary tieback
may be required
Angle with oversized hole
and plate washers

338

|

ARCHITECTURAL PRECAST CONCRETE

• If leveling bolt were recessed into
sill for ease of alignment, patching
might be required

Shim

DESIGN

4.5.10 Connection Details

Fig. 4.5.49 Shear plate (SP5).

Fig. 4.5.51 Shear plate (SP7).

Shown at bearing bracket

• (a) and (b) are sections at
horizontal joint. For vertical joint,
modify to eliminate overhead weld.

• A few of the variety of shear plates
used at bearing connectors

4

(a)

• (c) is section at vertical joint

• Shape and location of plate or
angle tailored to suit conditions
and forces to be resisted

• May be flush or recessed
• Shape and location of plates or
angles vary to suit conditions
and forces to be resisted
• Can also resist uplift
(b)

(c)

Fig. 4.5.52 Panel to panel alignment (PPA1).
• Not intended for required out-of-plane force resistance,
but can be adapted to serve as tieback, as in Fig. 4.5.68
• Dimension to face of panel is critical
• Good solution when slightly
bowed panels are not accessible
after erection

Fig. 4.5.50 Shear plate (SP6).
• Common at foundations

• If panels are accessible after
erection, finger plates can be
field welded and shimmed
if necessary

• May be flush or recessed
• Shape and location of plates or
angles vary to suit conditions and
forces to be resisted

Plan section
of variation

Fig. 4.5.53 Panel to panel alignment (PPA2).
• Not intended for required out-of-plane
force resistance, but can be adapted
to serve as tieback
• Shim thin panel if necessary

slotted
plate

ARCHITECTURAL PRECAST CONCRETE

|

339

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.54 Panel to panel alignment (PPA3).

Fig. 4.5.56 Column cover (CC1).

• Not intended for required out-of-plane force resistance,
but can be adapted to serve as tieback

• Serves as tieback

• Panels must be aligned before welding

• Length and diameter of rod may limit capacity

• Option (a) requires vertical weld, inside narrow joint

• First element of column cover must be aligned prior to placing
second

• Option (b) allows downhand weld, inside narrow joint

• Could be used for both halves if located at top
• Placement and
coverage of
insert is
difficult in
thin sections
• Shown in
conjunction
with Fig. 4.5.57

Threaded rod or coil
rod with nut and washer

Weld plate options
(a)

Inserts cast in panel

(b)
Joint
width

Fig. 4.5.55 Panel to panel alignment (PPA4).
• Not intended for required out-of-plane force resistance
• A few of the variety of alignment connectors
• Shape and location of embeds
and loose slugs tailored to
suit conditions
Threaded rod or coil
rod with nut and washer
Inserts cast in panel

Fig. 4.5.57 Column cover (CC2).
• Use where access is limited
• Exercise caution to prevent weld stain and
cracking from excess heat
• Minimum
recommended
joint size is
3/4 in. (19mm)

Plate cast in
column cover

Bar or plate
to match size
of joint

340

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.5.10 Connection Details

Fig. 4.5.58 Column cover (CC3).

Fig. 4.5.61 Column cover (CC6).

• Serves as tieback

• Bottom connector

• Used only at top for welding access

• Support unit on shims

• Could be changed to bolted

• Joint width sets thickness of vertical plate on knife assembly

4

• Align and weld first unit prior to setting second
Typical

• Welding of second half difficult in narrow joint
• Allows units to rock if bent plate (or angle) legs long enough

A

Knife assembly

Fig. 4.5.59 Column cover (CC4).
• Can be both a loadbearing and tieback connector
• Lower panel must be aligned and
welded prior to placing upper portion

Section A

• Good with limited access
• Could modify to insert and
bolt if ample space
Plan View

Fig. 4.5.62 Beam cover (BC1).
• Erection sequence critical
• Beam must be adequate to prevent excessive rotation
when first element is placed
• Top right connector (alternate) requires
tight tolerance

Section View

• S ealant at top left connector
(alternate) is critical
•M
 ay require combination of
grouting, bolting, and welding

Fig. 4.5.60 Column cover (CC5).
• Top connector

• P referably, use one type of
alternate top connector

• Align with tieback rods prior to welding angle

• J oint locations optional

Place 3rd

See DB2
See EB1

Place
2nd

Place 1st
Plan View

ARCHITECTURAL PRECAST CONCRETE

|

341

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.63 Soffit hanger (SH1).

Fig. 4.5.65 Special conditions (SC2).

• Allows alignment after in place

Concrete Anchors

• May require separate tiebacks for lateral forces

(a) National Coarse or coil thread loop insert

• Access for bolting
may be difficult

Oversized holes with
plate washers and nuts
(b) National Coarse or coil thread wing nut

(c) National Coarse or coil thread coupling nut
and bolt
TS
Threaded Rods
(d) National Coarse or coil thread coupling nut,
plate, and studs
Precast Concrete Soffit Panel

(e) Projecting National Coarse or coil bolt

Fig. 4.5.64 Special conditions (SC1).
Oversize hole considerations

(a) Bolt subject to bending

(b) L oose plate under angle, welded after
adjustment eliminates bending in bolt

342

|

ARCHITECTURAL PRECAST CONCRETE

(f) Flush plate with studs or hand welded
bolt blanks

(g) Bearing lug and/or tension bar supplement
on flush plate with studs or hand-welded
bolt blanks

(h) Proprietary threaded embedment.
Available with one- or two-way
adjustability.

DESIGN

4.5.10 Connection Details

Fig. 4.5.66 Special conditions (SC3).

Fig. 4.5.67 Special conditions (SC4).

Special tiebacks

Special shear plates that allow lift-off for rocking

4

(a) Use in pairs. Allows movement
perpendicular to panel

(a) Double pivots allow for extreme drift
Pivots

.
(b) Use in pairs. Round bars shop
welded to bearing bracket.

(c) Use in pairs. Round bar shop
welded to bearing bracket.

(b) S leeve eliminates possibility of binding when oversize hole
provides for drift
Sleeve

Fig. 4.5.68 Special conditions (SC5).
• Figure 4.5.52 can be supplemented to become
a tieback
• Lower panel with insert and bolt
must be aligned and welded prior
to placing upper panel
• Limited bolt head weld length
could be mitigated by shop
welding a plate to it

(c) Small diameter tieback rods desired for flexing can be prevented
from buckling in compression with loose pipe sleeve

Fig. 4.5.69 Special conditions (SC6).
Anchor load control shown at bearing bracket.
• Upper bolts carry shear
• Lower bolts carry tension

ARCHITECTURAL PRECAST CONCRETE

|

343

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.70 Bearing wall to foundation (BWF1).

Fig. 4.5.72 Bearing wall to foundation (BWF3).

• Can be designed for shear and uplift

• Has shear, uplift, and moment capacity

• Could develop moment resistance
by placing a connection on each
side of wall

• Location and alignment of dowels critical
• Capacity can be increased with
confinement reinforcing

• Shim prior to drypack

• Dowels projecting from panel create
storing and shipping problems
• Requires bracing until grouted
• Grouting could be done after
alignment if injection tube used
2 Shim
Stacks/Panel

Panels grouted
into sleeve

• Could be inverted with sleeve
and injection tube in panel

Variations

Fig. 4.5.73 Bearing wall to foundation (BWF4).
• Can be designed for shear, uplift,
and nominal moment capacity
• Requires bracing until welded

Concrete
Slab on Grade

Fig. 4.5.74 Bearing wall to foundation (BWF5).
• Bar could be prestressed or mild steel

Fig. 4.5.71 Bearing wall to foundation (BWF2).
• Insert must be jigged plumb
• Allows vertical adjustment without crane
• Finish joint with drypack or sealant
• Bolt head may be welded for tensile or shear capacity
• Plate may be eliminated, but adjustment becomes
more difficult
• May be inverted with insert below

• Substantial shear, uplift, and moment capacity
• Tolerance for placement of bars
and sleeves critical
• May require grout tubes
and vents
• Preferably grout from bottom
to eliminate voids

|

ARCHITECTURAL PRECAST CONCRETE

PostTensioning
Bar

• Bracing required until
drypacked and grouted
• Void in foundation at bar
essential for field alignment
• Foundation void can be formed
with EPS (expanded polystyrene) or
foam pipe insulation

344

Coupler

2 Shim Stacks/Panel

DESIGN

4.5.10 Connection Details

Fig. 4.5.75 Slab to bearing wall (SBW1).

Fig. 4.5.76 Slab to bearing wall (SBW2).

• Welding at bottom of tee
or slab is NOT recommended
as excessive restraint results

• Pocket and tee end must be planned so slab can be swung into
place

Coil Insert and Field
Placed Rod

• CANNOT be used at both ends of slab

• Load is eccentric
to wall panel
• Top connection for
shear can provide
some torsional
restraint of spandrel

4

• Consider volume change shortening of slab
• Pocket may telegraph through and show on outside of wall
(a)

• If slab at top of wall, as in (b), pockets could be replaced with
continuous dap

Variations

• Corbel requires
special forming
• Could replace
corbel with
inverted EB type
assembly

• DO NOT drypack pocket so it restricts tee stem

(b)

(a)

or

• Variations
(d) through (g)
could be used either
in topping or
hollow-core joints
(c)
Slotted Insert
Variation at joint
(d)
(b)

Reinforcement or
threaded insert
(dowel)

21/2 in. min.

(e)

Fig. 4.5.77 Slab to bearing wall (SBW3).
• DO NOT use at both ends of slab to prevent excessive restraint
• Develops a rigid moment connection

(f)

• Effect of moment, rotation, and volume
changes in wall and slab must be
considered
• Welding must be completed before
placing upper panel
• Avoid overhead welding if possible
• Could use wall corbel in lieu of
angle seat

(g)

Pre-Welded

ARCHITECTURAL PRECAST CONCRETE

|

345

4

DESIGN

4.5.10 Connection Details

Fig. 4.5.78 Slab to side wall (SSW1).

Fig. 4.5.79 Wall to wall (WW1).

• Allows for slab deflection

• Can be used to withstand uplift forces

• Transfers horizontal forces

• Connection is hidden and protected

• Do not over tighten bolt in (a)

• Connection is not developed until tensioning is completed (bars
are anchored)

• Proprietary or fabricated slot embedment in (b)
• Vertical movement accommodated by flexing plate
and welds in (c)
• Vertical movement accommodated by flexing tee flange in (d)
• (c) and (d) could be underneath, for less roofing interference, but
field labor would be more expensive

• Temporary bracing is required
• Drypack, tensioning, grouting sequence may limit erection to
one story at a time
• Grouting requires care to ensure complete filling

Threaded insert in panel
(a)
Fill Sleeve with Grout
Prior to Setting Bar
in Place
Shim and Grout

Lap Bar Grouted in Sleeve

(b)
Variations

Spiral Duct Sheathing
Threadbar Coupler

Shim & Drypack
Gasket at Sleeve
(c)

Grout Not Shown for Clarity
Spiral Duct Sheathing
Threadbar Coupler
Threadbar
Grout Tube w/nipple
Grout Tube Plug
Anchor Nut Pocket
Anchor Plate w/Vertical
Grout Vent Hole

(d)

Shim & Drypack
Grout Not Shown for Clarity

346

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.6 Tolerances / 4.6.2 Product Tolerances

4.6 T OLERANCES
4.6.1 G
 eneral
Designers must recognize that manufacturing and
erection tolerances apply to precast concrete just as
they do to other building materials. Tolerance is defined as a permissible variation from a specified dimension. A tolerance can be expressed as an additive (+)
or subtractive (-) variation from a specified dimension
or relation or as an absolute deviation from a specified
relation. Tolerances define realistic limits for size and
shape within which the precast concrete units must
lie, and must satisfy the designer’s intent while ensuring the constructability and economy of the building
system.
Three groups of tolerances should be established as
part of the precast concrete design: product tolerances, erection tolerances, and interfacing tolerances.
Tolerances are established for the following reasons:
1. S
 tructural—To ensure that structural design accounts for variations in dimensional control.
Examples include eccentric loading conditions,
bearing locations, hardware anchorage locations,
and locations of reinforcing or prestressing steel.
2. F easibility—To ensure acceptable performance
of joints and interfacing materials, such as glazing
between panels, and to ensure that designs and
details are dimensionally feasible.
3. V
 isual—To ensure that the variations will result in
an aesthetically acceptable structure.
4. E
 conomic—To ensure ease and speed of production and erection with a known degree of accuracy
in the dimensions of the precast concrete product.
5. C
 ontractual—To establish a known acceptability
range and to establish responsibility for developing, achieving, and maintaining mutually agreed
tolerances.
6. L egal—To avoid encroaching on property lines and
to establish tolerance standards against which the
work can be compared in the event of a dispute.
Tolerances and interface conditions are best handled
by the design team, GC, or other entity having the
contractual authority necessary to specify, coordinate,
and control interfacing requirements of other trades
that adjoin the precast concrete construction.
While the responsibility for specifying and maintain-

4

ing tolerances of the various elements may vary among
projects, it is important that this responsibility be clearly
assigned. The tolerances must be reasonable, realistic,
and within generally accepted limits. Some manufacturing and erection costs are directly proportional to
the tolerance requirements. It is more economical to
design connection and interace details with maximum
flexibility and to keep tolerance requirements as realistic as possible.
Whenever possible, it is preferred that product and
erection tolerance be specified in accordance with PCIrecommended values. PCI has published a comprehensive guide to tolerances in Tolerance Manual for Precast
and Prestressed Concrete Construction, MNL 135.
It should be understood by those involved in the design and construction process that the listed tolerances
in this manual must be considered as guidelines for an
acceptability range and not limits for rejection. If specified tolerances are met, the members should be accepted. If these tolerances are exceeded, the member
may still be acceptable if it meets any of the following
criteria:
1. Exceeding the tolerances does not affect the structural integrity, architectural performance of the
member, or other trades.
2. The member can be brought within tolerance by
structurally and architecturally satisfactory means.
3. The total erected assembly can be modified reasonably to meet all structural and architectural
requirements.
The enforcement of tolerances should be based on
the technical judgment of the designer. This design
professional is able to decide whether a deviation from
the allowable tolerances affects safety, appearance, or
other trades. In building construction, very little out of
tolerance work, whether it is concrete, masonry, castin-place concrete, steel, or precast concrete, has been
rejected and removed solely because it was “out-oftolerance.”

4.6.2 Product Tolerances
Product tolerances relate to the dimensions and dimensional relationships of the individual precast concrete units. They are a measure of the dimensional
accuracy required on individual members to ensure,
prior to delivery, that the members will fit the structure
without requiring tolerance related rework. Product

ARCHITECTURAL PRECAST CONCRETE

|

347

4

DESIGN

4.6.2 Product Tolerances

tolerances are applied to physical dimensions of units
such as thickness, length, width, squareness, and location and size of openings. They are determined by economical and practical production considerations, and
functional and appearance requirements. Product tolerances also control the locations of the member features as they relate to the overall member dimensions.
The architect should specify product tolerances within generally accepted limits, as they relate to each individual project, or require performance within a generally accepted limit. The architect must account for the
function of the member, its fit in the structure, and
the compatibility of the member tolerances to those of
the interfacing materials (see Section 4.6.4). Tolerances
for manufacturing are standardized throughout the industry and should not be specified to more stringent,
and therefore more costly, unless absolutely necessary.
Areas that might require more exacting tolerances
could include special finish or appearance requirements, glazing details, and certain critical dimensions
on open shaped panels (see Sections 3.5, 5.2. and
3.3.1).
For example, a special appearance requirement may
be necessary for honed or polished flat concrete walls
where bowing or warping tolerances may have to be
decreased to avoid joint shadows. Another special
case might be tolerances for dimensions controlling
the matching of open-shaped panels. These tolerances
may have to be tighter than the standard dimensional
tolerances to ensure a visually acceptable match-up,
unless the architect has recognized and solved the
alignment problem as part of the design.
When a project involves particular features sensitive to
the cumulative effect of generally accepted tolerances
on individual portions, the design team should anticipate and provide for this effect by setting a cumulative tolerance or by providing escape areas (clearances)
where accumulated tolerances or production errors
can be absorbed. The consequences of all tolerances
permitted on a particular project should be investigated to determine whether a change is necessary in the
design or in the tolerances applicable to the project or
individual components. For example, there should be
no possibility of minus tolerances accumulating so that
the bearing length of members is reduced below the
required design minimum. These bearing dimensions
and their tolerances should be shown on the erection
drawings.

348

|

ARCHITECTURAL PRECAST CONCRETE

The published allowable variation for one element
of the structure should not be applicable when it will
permit another element of the structure to exceed its
allowable variations.
Restrictive tolerances should be reviewed to ascertain
that they are compatible and that the restrictions can
be met. For example, a requirement that states, “no
bowing, warpage, or movement is permitted,” is not
practical or possible to achieve.
The product tolerances for architectural precast concrete panels have the following significance:
1. Length or width dimensions and straightness of the
precast concrete will affect the joint dimensions,
opening dimensions between panels, and possibly
the overall length of the structure. Tolerances must
relate to unit size and increase as unit dimensions
increase.
2. Panels out-of-square can cause tapered joints and
make adjustment of adjacent panels difficult.
3. Thickness variation of the precast concrete unit becomes critical when interior surfaces are exposed
to view. A non-uniform thickness of adjacent panels will cause offsets of the front or the rear faces
of the panels.
Industry product tolerances for architectural precast concrete panels are defined as follows:
Warping is generally the twisting of a member, resulting in an overall out-of-plane curvature in which
the corners of the panel do not all fall within the same
plane. Warping tolerances are stated in terms of the
magnitude of the corner variation (Fig. 4.6.1). It is usually stated in terms of the allowable variation per 1
ft (0.3 m) of distance from the nearest adjacent corner with a not-to-exceed maximum value of corner
warping.
Bowing is an overall out-of-plane curvature, which
differs from warping in that while the corners of the
panel may fall in the same plane, the portion of the
panel between two parallel edges is out of plane.
Several possible bowing conditions are shown in Fig.
4.6.2. Differential temperature effects, differential
moisture absorption between the inside and outside
faces of a panel, the effects of prestress eccentricity,
and differential shrinkage between wythes in an insulated panel should be considered in design to minimize
bowing and warping.

DESIGN

4.6.2 Product Tolerances

Fig. 4.6.1 Warping definitions for panels.

4

Fig. 4.6.2 Definition of bowing for panels.
Length of bow

Length of bow
Exposed face

True Plane

Distance to farthest
adjacent corner

D
ad ista
jac nc
en e to
tc n
or ea
ne re
r st

Corner
Warping
Max. bowing (j)
Cross section convex bowing

Variation in local
smoothness

Cross section concave bowing

Bowing (cross section)

Fig. 4.6.3 Local smoothness variations.

10' (3 m)

Max. bowing (j)

Exposed
surface of
precast
concrete
panel

Exposed face (convex)

/8” (9 mm) shim (typ.)

3

1
/4” (6 mm) roller (typ.)
(fits anywhere)

10' (3 m) straightedge (typ.)

Length of bow

Exposed face (concave)

Max. bowing (j)

Variation in local
smoothness

/8” (9 mm) shim (typ.)

3

Bowing (j)
(elevation)
Length of
bow
(elevation)

Precast
concrete
panel

1
/2” (13 mm) roller (typ.)
(won’t fit anywhere)

Measuring local smoothness
variations of any surface
Elevation

Bowing and warping tolerances are of interest primarily at the time the panel is erected. They have an
important effect on the edge match-up during erection and on the visual appearance of the erected panels, both individually and when viewed together. The
requirements for bowing and warping of panels may
be overridden by erection tolerances for panels as installed with reference to joint widths, jog in alignment,
and step-in face.
Tolerances for the planeness of concrete surfaces at
the window or curtain wall face should be provided.
Bowed or warped panels can make alignment of adja-

Length of bow
(cross section)
Panel bowed in both elevation &
cross section

cent panels or materials difficult. Slender panels should
not be automatically subjected to the standard tolerances for bowing and warping. Table 4.6.1 shows the
relationship between overall flat panel dimensions and
thickness. Anything below the suggested bowing and
warping tolerances should be reviewed to determine
if the dimensions should possibly be increased or a
mid-point tieback used. Note that the thickness values in this table should not be considered as limiting
values, but rather as an indication that more detailed
consideration of the possible magnitude of warping
and bowing is warranted. The major criteria for maintaining or relaxing bowing and warping tolerances will

ARCHITECTURAL PRECAST CONCRETE

|

349

4

DESIGN

4.6.2 Product Tolerances / 4.6.3 Erection Tolerances

Table 4.6.1. Guidelines for Panel Thickness for Overall Panel Stiffness Consistent with
Suggested Normal Panel Bowing and Warping Tolerances.1

Panel
Dimensions2

8 ft

10 ft 12 ft 16 ft 20 ft 24 ft 28 ft 32 ft

4 ft

4 in.

4 in.

4 in.

5 in.

5 in.

6 in.

6 in.

7 in.

6 ft

4 in.

4 in.

4 in.

5 in.

6 in.

6 in.

6 in.

7 in.

8 ft

4 in.

5 in.

5 in.

6 in.

6 in.

7 in.

7 in.

8 in.

10 ft

5 in.

5 in.

6 in.

6 in.

7 in.

7 in.

8 in.

8 in.

1

This table should not be used for panel thickness selection.

2

T his table represents a relationship between overall flat panel dimensions and thickness below which
suggested bowing and warping tolerances should be reviewed and possibly increased or a mid-point
tieback used. For ribbed panels, the equivalent thickness should be the overall thickness of such ribs if
continuous from one end of the panel to the other.

Note: 1 ft = 0.3048 m; 1 in. = 25.4 mm

be the appearance requirements, the required type of
connections (as well as the number and location of tieback connection points), and the advice of the local
precaster regarding overall economic and construction
feasibility.
To reduce the possibility of panel warpage or bowing, consideration should be given to the panel length,
shape, and connection locations. The longer the panel,
the more difficult it is to control planeness of the panel. Bowing or warpage can be reduced by the use of
multiple tieback connections.
For ribbed panels, the equivalent thickness used in
Table 4.6.1 should be the overall thickness of the ribs,
if they run continuous from one end of the panel to
the other. Similarly, panels that are manufactured using large-aggregate concrete (above 3/4 in. [19 mm]
aggregate) or units that are fabricated from nonhomogeneous materials (such as two significantly different concrete mixtures, natural stone or clay product
veneers, insulating mediums, and the like) also require
more careful consideration of all aspects of fabrication, storage, and handling with regard to bowing and
warping.
Surface out-of-planeness is defined as a local
smoothness variation rather than a bowing of the entire panel shape. Examples of local smoothness variations are shown in Fig. 4.6.3. The tolerance for this
type of variation is usually expressed in fractions of 1
in. per 10 ft (25mm per 3 m).
Figure 4.6.3 also shows how to determine if a surface
meets a tolerance of 1/4 in. per 10 ft (6 mm per 3 m).
A 1/4 in. (6 mm) diameter by 2-in.-long (50 mm) roll-

350

|

ARCHITECTURAL PRECAST CONCRETE

er should fit anywhere between the 10-ft-long (3 m)
straightedge and the element surface being measured
when the straightedge is supported at its ends on 3/8
in. (10 mm) shims as shown. A 1/2 in. (13 mm) diameter
by 2-in.-long (50 mm) roller should not fit between the
surface and the straightedge.
Dimensional tolerance requirements for architectural precast concrete elements are given in Fig. 4.6.4,
4.6.5, 4.6.6, and 4.6.7. It must be emphasized that
these are guidelines only and that each project must be
considered individually to ensure that the staled tolerances are applicable.
Groups of inserts or cast-in items that must be located in close tolerance to each other should not be
separated into two different panels by a joint.

4.6.3 Erection Tolerances
Erection tolerances control the position of the individual precast concrete members as they are located
and placed in the assembled structure. They normally
involve the GC and various subcontractors, such as the
precast concrete erector.
Erection tolerances are provided to help achieve uniform joint widths, level floor elevations, and planar
wall conditions. Erection tolerances should be determined on the basis of individual unit design, shape,
thickness, composition of materials, and overall scale
of the unit in relation to the building. The specified
erection tolerances may affect the work of several different building trades and must be consistent with the
tolerances specified for those trades.

DESIGN

4.6.3 Erection Tolerances

4

Fig. 4.6.4 Architectural Wall Panels.
Allowable angle of rotation l4
b1

b3
c1
b3

c2

c

b2

r1

c2
a2

b1
b4

Architecturally
exposed surface

a1
n2

h
o

n2

l1
l2

l1

o

t2

ace
kF

t2
c1

c

Ba

b8
n1

n1

i
b

a1

t1
c

e

c

a2

a1 =	Overall height of unit measured at the face exposed to view:
Up to 10 ft [3 m] ........................................................................................... ± 1/8 in. [± 3 mm]
10 to 20 ft [3 to 6 m] .......................................................... + 1/8 in., –3/16 in. [+ 3 mm, –5 mm]
20 to 40 ft [6 to 12 m] .................................................................................. ± 1/4 in. [± 6 mm]
Greater than 40 ft [12 m] . ............................................ ± 1/16 in. per 10 ft [± 1.5 mm, per 3 m]
a2 =	Overall height of unit measured at the face not exposed to view:†
Up to 10 ft [3 m] ........................................................................................... ± 1/4 in. [± 6 mm]
10 to 20 ft [3 to 6 m] ......................................................... + 1/4 in., –3/8 in. [+ 6 mm, –10 mm]
20 to 40 ft [6 to 12 m] ................................................................................ ± 3/8 in. [± 10 mm]
Greater than 40 ft [12 m] . ................................................ ± 1/8 in. per 10 ft [± 3 mm, per 3 m]
b

=	Overall width of unit measured at the face exposed to view:
Up to 10 ft [3 m] ........................................................................................... ± 1/8 in. [± 3 mm]
10 to 20 ft [3 to 6 m] .......................................................... + 1/8 in., –3/16 in. [+ 3 mm, –5 mm]
20 to 40 ft [6 to 12 m] .................................................................................. ± 1/4 in. [± 6 mm]
Greater than 40 ft [12 m] . ............................................. ± 1/16 in. per 10 ft [± 1.5 mm per 3 m]

b1 = Rib width ...................................................................................................... ± 1/8 in. [± 3 mm]
b2 = Distance between ribs . ................................................................................. ± 1/8 in. [± 3 mm]
b3 = Rib to edge of flange .................................................................................... ± 1/8 in. [± 3 mm]
b8 =	Overall width of unit measured at the face not exposed to view:
Up to 10 ft [3 m] ........................................................................................... ± 1/4 in. [± 6 mm]
10 to 20 ft [3 to 6 m] ......................................................... + 1/4 in., –3/8 in. [+ 6 mm, –10 mm]
20 to 40 ft [6 to 12 m] ................................................................................ ± 3/8 in. [± 10 mm]
Greater than 40 ft [12 m] . ................................................. ± 1/8 in. per 10 ft [± 3 mm per 3 m]
c

= Total thickness ..................................................................... + 1/4 in., –1/8 in. [+ 6 mm, –3 mm]

c1 = Flange thickness .................................................................. + 1/4 in., –1/8 in. [+ 6 mm, –3 mm]
c2 = Dimensions of haunches ............................................................................... ± 1/4 in. [± 6 mm]

ARCHITECTURAL PRECAST CONCRETE

|

351

4

DESIGN

4.6.3 Erection Tolerances

e

=	Variation‡ from square or designated skew . .............................. ± 1/8 in. per 6 ft, ± 1/2 in. max.
[± 3 mm per 2 m, ± 13 mm max.]

h

= Location smoothness, unconcealed surfaces .................... ± 1/4 in. per 10 ft, [± 6 mm per 3 m]

i

= Bowing .............................................................. ± Length/360, to maximum of 1 in. [25 mm]

j

= Warp (from adjacent corner) .................................................. 1/16 in. per ft [1.5 m per 300 mm]

l1 = Location of weld plates . ............................................................................... ± 1 in. [± 25 mm]
l2 = Tipping and flushness of plates . ................................................................... ± 1/4 in. [± 6 mm]
l4 =	Allowable rotation of plate, channel insert, electrical box .......................................... 2 degrees
1 4
/ in. [6 mm] maximum measured at perimeter of insert
m2 = Haunch bearing surface tipping and flushness of bearing plates .................. ± 1/8 in. [± 3 mm]
m3 =	Difference in relative position of adjacent haunch bearing surfaces
from specified relative position . ................................................................... ± 1/4 in. [± 6 mm]
n1 = Location of opening within panel ................................................................. ± 1/4 in. [± 6 mm]
n2 = Length and width of blockouts and openings within one unit ...................... ± 1/4 in. [± 6 mm]
n3 =	Location and dimensions of blockouts hidden from view
and used for HVAC and utility penetrations ................................................ ± 3/4 in. [± 19 mm]
o

= Position of sleeve ......................................................................................... ± 1/2 in. [± 13 mm]

p

= Reinforcing steel extending out of member ................................................. ± 1/2 in. [± 13 mm]

q

= Position of handling devices ......................................................................... ± 3 in. [± 75 mm]

r1 = Location of bearing surface from end of member ......................................... ± 1/4 in. [± 6 mm]
s1 =	Reinforcing steel and welded wire reinforcement:
Where position has structural implications or affects concrete cover ........... ± 1/4 in. [± 6 mm]
Otherwise .................................................................................................... ± 1/2 in. [± 13 mm]
s3 = Position of insert .......................................................................................... ± 1/2 in. [± 13 mm]
s4 =	Location of strand:
Perpendicular to panel .................................................................................. ± 1/4 in. [± 6 mm]
Parallel to panel ............................................................................................ ± 1 in. [± 25 mm]
t1 = Dimensions of architectural features and rustications .................................. ± 1/8 in. [± 3 mm]
t2 = Location of rustication joints . ....................................................................... ± 1/8 in. [± 3 mm]
w1 = Location of flashing reglets . ......................................................................... ± 1/4 in. [± 6 mm]
w2 = Location of flashing reglets at edge of panel ................................................ ± 1/8 in. [± 3 mm]
w3 = Size of reglets for glazing gaskets . ............................................................... ± 1/8 in. [± 3 mm]
z

= Electrical outlets, hose bibs, etc. . ................................................................. ± 1/2 in. [± 13 mm]

Tolerances below are for smooth-finished stone veneer-faced precast concrete panels.
1. Variations in cross-sectional dimensions: For thickness of walls from dimensions indicated..............
. ....................................................................................................................................1/4 in. (6 mm).
2. Variation in joint width: 1/8 in. in 36 in. (3 mm in 900 mm) or a quarter of nominal joint width,
whichever is less.
3. Variation in plane between adjacent stone units (lipping): 1/16 in. (1.5 mm) difference between
planes of adjacent units.
*Units shall be manufactured so that the face of each unit which is exposed to view after erection
complies with the following dimensional requirements.
†Unless joint width and fit-up requirements demand more stringent tolerance.
‡Applies to both panel and to major openings in panel. Tolerances apply to the difference of the two
diagonal measurements.

352

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.6.3 Erection Tolerances

4

Fig. 4.6.5 Brick faced architectural elements.

d
A

e

b

a

A-A
e

c

b

a

A

d

a

=	Alignment of mortar joints:
Jog in alignment ............................................................................................... 1/8 in. [± 3 mm]
Alignment with panel centerline ................................................................... ± 1/8 in. [± 3 mm]

b

= Variation in width of exposed mortar joints . ................................................ ± 1/8 in. [± 9 mm]

c

=	Tipping of individual bricks from the panel plane of exposed brick surface
–1/4 in. [–6 mm]
≤ depth of form liner joint

d

=	Exposed brick surface parallel to primary control surface of panel ........................ + 1/4 in., –1/8
[+ 6 mm, –3 mm]

e

=	Individual brick step in face from the panel plane of exposed brick surface .... –1/4 in. [–6 mm]
≤ depth of form liner joint

Note: The number of bricks that could exhibit these misalignments should be limited to 2% of the
bricks on the panel. See other panel tolerances in Fig. 4.6.4.

ARCHITECTURAL PRECAST CONCRETE

|

353

4

DESIGN

4.6.3 Erection Tolerances

Fig. 4.6.6 Columns.
a
r1
q2

b

c

r2

k

I1

Cross Section

z

h

p
10'

e

e

I1
p
Plan
d

r2

I2
q2

f
Elevation

a

= Length . ............................................................................................................................................................................... ± 1/2 in. [± 13 mm]

b

= Width .................................................................................................................................................................................... ± 1/4 in. [± 6 mm]

c

= Depth .................................................................................................................................................................................... ± 1/4 in. [± 6 mm]

d

= Variation from specified plan and squareness or skew ..................... ± 1/8 in. per 12 in., ± 3/8 in. max. [± 3 mm per 300 mm, ± 10 mm max.]

e

= Variation from specified elevation end squareness or skew .............. ± 1/8 in. per 12 in., ± 3/8 in. max. [± 3 mm per 300 mm, ± 10 mm max.]

f

= Sweep ....................................................................................................... ± 1/8 in. per 10 ft, ± 1/2 in. max. [± 3 mm per 3 m, ± 13 mm max.]

h

= Local smoothness of any surface ....................................................................................................................... 1/4 in. per 10 ft [6 mm in 3 m]

k

= Location of strand ................................................................................................................................................................ ± 1/4 in. [± 6 mm]

l1 = Location of embedment ....................................................................................................................................................... ± 1 in. [± 25 mm]
l2 = Tipping and flushness of embedment ................................................................................................................................... ± 1/4 in. [± 6 mm]
p

= Location of inserts for structural connections ..................................................................................................................... ± 1/2 in. [± 13 mm]

q1 = Location of handling device parallel to length of member ................................................................................................. ± 6 in. [± 150 mm]
q2 = Location of handling device transverse to length of member .............................................................................................. ± 1 in. [± 25 mm]
r1 = Location of haunch bearing elevation from end . .................................................................................................................. ± 1/4 in. [± 6 mm]
r2 = Variation from specified haunch bearing surface slope . .................... ± 1/8 in. per 12 in., ± 3/8 in. max. [± 3 mm per 300 mm, ± 10 mm max.]
z

354

|

= Base plate overall dimensions . ............................................................................................................................................. ± 1/4 in. [± 6 mm]

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.6.3 Erection Tolerances

4

Fig. 4.6.7 Architectural trim units.
Sills, Lintels, Copings, Cornices, Quoins, and Medallions
Bollards, Benches, and Planters
c

c1
t

a

= Height or length . .................................. ± 1/4 in. [± 6 mm]

b

= Width or diameter . ............................... ± 1/4 in. [± 6 mm]

o

=	Location of inserts and appurtenances:
Formed surfaces .................................... ± 1/4 in. [± 6 mm]
Unformed surfaces ................................ ± 1/4 in. [± 6 mm]

t

= Size / location of rustication / features .....± 1/4 in. [± 6 mm]

j

b1
b

b
i

h
a
a

=	Length . ............................................................................................. ± 1/8 in. [± 3 mm]
Where one face will be installed in dead wall space of mortar joint . ± 1/4 in. [± 6 mm]

b

= Overall width of units* ..................................................................... ± 1/8 in. [± 3 mm]

b1 =	Location of inserts and appurtenances:
On formed surfaces . ......................................................................... ± 1/8 in. [± 3 mm]
On unformed surfaces . ..................................................................... ± 3/8 in. [± 9 mm]
c

a
t
o

= Overall height of units* .................................................................... ± / in. [± 3 mm]
1 8

c1 =	Total thickness ................................................................................. ± 1/8 in. [± 3 mm]
Flange thickness .............................................................................. ± 1/8 in. [± 3 mm]
Where one face will be installed in dead wall space of mortar joint . ± 1/4 in. [± 6 mm]
t

= Size and location of rustications and architectural features . ........ ± 1/16 in. [± 1.5 mm]

h

= Local smoothness . ................................................ ± 1/8 in. per 5 ft [± 3 mm per 1.5 m]

i

= Bowing ................................................................... span / 360, max., ± 1/4 in. [± 6 mm]

j

= Warping† ............................................................ ± 1/16 in. per ft. [± 1.5 mm per 0.3 m]

b

Pavers
a

*Measured at face exposed to view.
†Measured per foot of distance from nearest adjacent corner.

a
Pavers
a

= Length or width ............................... ± 1/16 in. [± 1.5 mm]

c

= Thickness ......................................... ± 1/16 in. [± 1.5 mm]

j

= Warping* ....................................... ± 1/32 in. [± 0.75 mm]

*Measured per foot [0.3 m] of distance from nearest adjacent corner.
c
j

ARCHITECTURAL PRECAST CONCRETE

|

355

4

DESIGN

4.6.3 Erection Tolerances

Erection tolerances are both equipment and site dependent. There may be valid reasons to vary some of
the recommended tolerances to account for unique
project conditions. The erection tolerances should be
carefully reviewed by the designer and the involved
contractors and adjusted, if necessary, to meet the
project requirements. The effects of adjusted tolerances on specific details at joints, connections, and in other locations in the structure should be evaluated by the
designer. Different details may have varying amounts
of sensitivity to tolerances. If the final erection tolerances are different from those given in this manual,
the tolerances should be stated in writing and noted
on the project erection drawings.

ondary control surfaces of a precast concrete member
may be directly additive, the erection drawings should
clearly define the primary erection control surfaces. If
both primary and secondary control surfaces are critical, provisions for adjustment should be included. The
accumulated tolerance limits may be required to be
accommodated in the interface clearance. This may
occur with window openings between two spandrels
when the critical elevation, top or bottom and as indicated on the erection drawings, must be maintained.
If more than one critical line is indicated, the erector
should balance any deviations between the two edges.
Surface and feature control requirements should be
clearly outlined in the plans and specifications.

The erector is responsible for erecting the members
within the specified tolerances and completing the
connections in the manner specified. Appropriate
surveying and layout procedures should be followed
to ensure accurate application of tolerances. When a
unit cannot be erected within the specified tolerances,
the erector should notify the precaster and GC/CM to
check the structural adequacy of the installation and
determine if the connection design should be modified. No unit should be left in an unsafe condition. Any
adjustments affecting structural performance, other
than adjustments within the prescribed tolerances,
should be made only after approval by the precast concrete design engineer.

During wall panel installation, priority is generally
given to aligning the exterior face of the units to meet
aesthetic requirements. This may result in the interior
face of units being out-of-plane.

The primary control surfaces or features on the precast concrete members are erected to be in conformance with the established erection and interfacing
tolerances (Figs. 4.6.10 and 4.6.11). Clearances are
generally allowed to vary so that the primary control
surface can be set within tolerance. It is important to
recognize product tolerances are not additive to the
primary surface erection tolerances.
Secondary control surfaces that are positioned from the
primary control surfaces by the product tolerances are
usually not directly positioned during the erection process, but are controlled by the product tolerances. Thus,
if the primary control surfaces are within erection and
interfacing tolerances, and the secondary surfaces are
within product tolerances, the member should be considered erected within tolerance. The result is that the
tolerance limit for secondary surfaces may be the sum of
the product and erection tolerances. Product tolerances,
in general, must not exceed erection tolerances.
Because erection and product tolerances for some sec-

356

|

ARCHITECTURAL PRECAST CONCRETE

Erection tolerances are largely determined by the actual alignment and dimensional accuracy of the building foundation and frame in those circumstances where
the building frame is constructed from some material
other than precast concrete. The GC is responsible for
the plumbness, levelness, and alignment of the foundation and non-precast concrete structural frame, including the location of all bearing surfaces and anchorage points for precast concrete products.
The architect/engineer should clearly define in the
specifications the maximum tolerances permitted in
the foundation and building frame alignment, then
should specify that the GC check frequently to verify
these tolerances are being held. In addition, the architect/engineer should ensure that the details in the contract documents allow for the specified tolerances. To
accommodate any misalignment of the building frame,
connections should provide for vertical, horizontal, and
lateral adjustments of at least 1 in. (25 mm).
If the precast concrete units are to be installed reasonably “plumb, level, square, and true,” the actual
location of all surfaces affecting their alignment, including the levels of floor slabs and beams, the vertical
alignment of floor slab edges, and the plumbness of
columns or walls, must be known before erection begins. The GC is expected to, and should be required to,
establish (and maintain at convenient locations) control points, benchmarks, and lines in areas that remain
undisturbed until the completion and acceptance of
the project.

DESIGN

4.6.3 Erection Tolerances

Tolerances for the building frame must be adequate
to prevent interferences that may cause difficulty with
panel installation. Whenever possible, beam elevations
and column locations should be uniform in relation to
the precast concrete units with a constant clear distance between the precast concrete and the support
elements.
The location of hardware items cast into or fastened
to the structure by the GC, steel fabricator, or other
trades, should be located ±1 in. (±25 mm) in all directions (vertical and horizontal) from the specified location, plus a slope deviation of no more than ±1/4 in. (±6
mm) in 12 in. (300 mm) for critical bearing surfaces.
Connection details, therefore, should consider the possibility of bearing surfaces being misaligned or warped
from the desired plane, which would necessitate approved field adjustments to be made.
In the determination of erection tolerances, attention should also be given to possible deflection and/
or rotation of structural members supporting precast
concrete. This is particularly important when bearing on flexible or cantilevered structural members.
Consideration should be given to both initial and to
long-term deflection caused by creep of the supporting structural members. Specific tolerances cannot be
assigned to erection deformations. By considering realistic tolerance variations and clearances, the influence
of erection deformations can be minimized or eliminated for practical purposes.
Structural steel framing tolerances should be
specified to conform with the American Institute of
Steel Construction (AISC) Code of Standard Practice
for Steel Buildings and Bridges. Particular attention is
directed to the “Commentary” included in this code,
which provides a detailed explanation of the specified
erection tolerances. Mill, fabrication, and erection tolerances combined result in the final dimensional accuracy of the structural steel frame.
Precast concrete wall tolerances should follow those
for the steel frame, because the allowable tolerances
for steel frame structures make it impractical to maintain precast concrete panels in a true vertical plane in
tall structures. Based on the allowable steel frame tolerances, it would be necessary to provide for a 3 in. (75
mm) adjustment in connections up to the 20th story
(Fig. 4.6.8). Above the 20th story, the façade may be
maintained within 1/16 in. (1.6 mm) per story with a
maximum total deviation of 1 in. (25 mm) from a true

4

Fig. 4.6.8 Clearance example.

Theoretical CL of Column
Max. Position Out
“Toward” Bldg. Line
Possible Position of
Precast Concrete
Façade

2"
3"

Max. Position in “Away” from Bldg. Line

/"

1 4

Column Out
Column In

36 Stories
Steel Framing Constructed Per
AISC Code of Standard Practice

1"

20 Stories
2"

Theoretical Plane
of Precast Concrete
Façade “Building
Line”

Note: 1 in. = 25.4 mm.

vertical plane, if connections that provide for 3 in. (75
mm) of adjustment are used. Connections that permit
adjustments of +2 in. (+50 mm) to -3 in. (-75 mm) (5
in. [125 mm] total) will be necessary in cases where it is
desired to construct the façade to a true vertical plane
above the 20th story. These adjustments in connections are not economically feasible.
A solution that has proven both practical and economical is to specify the more stringent AISC elevator-column erection tolerances for steel columns in the
building façade that will receive the precast concrete
panels. This type of solution should be agreed to as part
of the design and specification process, or at least prior
to finalization of the fabrication erection process.
Cast-in-place concrete frame tolerances are given in ACI 117, Standard Tolerances for Concrete
Construction and Materials, unless otherwise stated in
the contract documents. ACI tolerances are not realistic for tall buildings. Also, greater variations in height
between floors are more prevalent in cast-in-place con-

ARCHITECTURAL PRECAST CONCRETE

|

357

4

DESIGN

4.6.3 Erection Tolerances

crete structures than in other structural frames. This
may affect the location or mating of the inserts in the
precast concrete units with the cast-in connection devices. Tolerances for cast-in-place concrete structures
may have to be increased further to reflect local trade
practices, the complexity of the structure, and climatic
conditions. As a result, it is recommended that precast
concrete walls should follow concrete frames in the
same manner as for steel frames, if the details allow it
and appearance is not affected.

b. Variation in elevation from lines parallel to
specified grade lines:
• 1 /40 in./ft (0.7 mm/0.3 m) for adjacent members less than 20 ft (6 m) apart or any wall or
bay length less than 20 ft.
• 1/2 in. (13 mm) for adjacent members 20 ft (6
m) or more apart or any wall or bay length of
20 ft or more.
3. Anchor bolts
a. Variation from specified location in plan:

The following tolerances, in addition to ACI 117
requirements, should be specified for cast-in-place
concrete to which precast concrete units are to be
connected:

• 3 /4 in. (19 mm) and 7/8 in. (22 mm) bolts ......
± 1/4 in. (6 mm).
• 1 in. (25 mm), 11/4 in. (32 mm), and 11/2 in.
(38 mm) bolts ......± 3/8 in. (10 mm).

1. F ootings, caisson caps, and pile caps

• 13/4 in. (44 mm), 2 in. (51 mm), and 21/2 in.
(64 mm) bolts ......± 1/2 in. (13 mm).

a. V
 ariation of bearing of surface from specified
elevation: ±1/2 in. (±13 mm)

b. Variation center to center of any two bolts within an anchor bolt group: ≤ 1/8 in. (≤3 mm).

2. Piers, columns, and walls
a. Variation in plan from straight lines parallel to
specified linear building lines:

c. Variation from specified elevation: ± 1/2 in. (± 13
mm).

• 1 /40 in./ft (0.7 mm/0.3 m) for adjacent members
less than 20 ft (6 m) apart or any wall or bay
length less than 20 ft.

d. Anchor bolt projection: – 1/4 in., + 1/2 in. (- 6 mm,
+ 13 mm).
e. Plumbness of anchor bolt projection: ± 1/16 in./ft
(± 1.6 mm/0.3 m)

• 1/2 in. (13 mm) for adjacent members 20 ft (6
m) or more apart or any wall or bay length of
20 ft or more.

AISC recommends hole sizes and position tolerance
for various bolt diameters as follows:

Table 4.6.2. Anchor rod hole diameter and postion tolerance.

358

|

ARCHITECTURAL PRECAST CONCRETE

Anchor Rod
Diameter, in. (mm)

Hole Diameter,
in. (mm)

Position Tolerance,
in. (mm)

3

/4 (19)

15/16 (33)

±1/4 (±6)

7

/8 (22)

1 /16 (40)

±1/4 (±6)

1 (25)

113/16 (46)

±3/8 (±10)

11/4 (32)

21/16 (52)

±3/8 (±10)

11/2 (38)

25/16 (59)

±3/8 (±10)

13/4 (44)

2¾ (70)

±1/2 (±13)

2 (50)

3¼ (82)

±1/2 (±13)

21/2 (63)

3¾ (95)

±1/2 (±13)

9

DESIGN

4.6.3 Erection Tolerances

4

Fig. 4.6.9 Design concepts to accommodate site tolerances.

Recommended

It should be recognized that ACI 117 applies primarily
to reinforced concrete buildings, and the AISC Code of
Standard Practice applies only to steel building frames.
Neither of these standards apply to buildings of composite construction (that is, concrete floor slabs supported by steel columns or concrete-encased structural
steel members, fireproofed frames, or steel frames with
precast concrete cladding). Obviously, the location of
the fireproofing face on the steel, as well as that of the
steel member itself, are both critical. Because the alignment of composite members, fireproofing, and masonry
work are not controlled by referencing these standards,
the architect/engineer should require that the location
of all such materials contiguous to the precast concrete
units be controlled within tolerances that are no less
stringent than those specified in ACI 117. Should there
be some doubt as to what these tolerances should be,
the precast concrete manufacturer or erector should be
consulted for advice.
It is generally poor practice to design gaps or joints
between site work and precast concrete as an architectural feature. A case in point would be the designing of
individual panels to fit between cast-in-place columns
and beams with either of these structural members exposed (Fig. 4.6.9). Unless the cast-in-place structure is
executed to well above normal tolerances, the width
of joints must be allowed with a large tolerance (± 1/2
in. [± 13 mm] in the case of a 20 ft [6 m] opening).
The actual joint width may differ in each bay, and will
certainly require sealants with corresponding flexibil-

Not Recommended

ity. Joint widths may be adjusted to enable them to
be equal at either end of a panel, but efforts toward
equalizing the joints on either side of a column cannot
be attempted unless panels can be adjusted horizontally after erection. The problems this could cause are
avoided where the cladding passes in front of the columns and the jointing is between the panel edges.
Total precast concrete systems—Erection tolerances are less critical in structures consisting entirely
of precast concrete units than for structures that are
combinations of precast and cast-in-place concrete
or steel. Where precast concrete units connect to site
work, such as at footings or foundation walls, larger
erection tolerances are particularly necessary.
Erection tolerances—The erection tolerances of
architectural precast concrete are given in Fig. 4.6.10
and 4.6.11. These are guidelines only and each project must be considered individually to ensure that the
stated tolerances are applicable. After precast concrete
erection and before other trades interface any materials with the precast concrete members, it should be
verified that the precast concrete elements are erected
within the specified tolerances.
Because a panel base connection often allows some
positioning flexibility, it is often more important to
control dimensions from haunch to haunch in walls or
multistory columns rather than to maintain tight control of actual haunch location dimensions from the end
of the member.

ARCHITECTURAL PRECAST CONCRETE

|

359

4

DESIGN

4.6.3 Erection Tolerances

Fig. 4.6.10 Architectural walls/spandrels erection tolerances.
Bldg. grid
datum

a

i

h

Horizontal primary
control surface

dd

f

a1

CL of steel
suport

b
c
Support
member

Side View
Walls

Plan View
Walls
Bldg. grid
datum

e

e

10 ft 0 in. (3m)

a

a

h10
g

Elevation View
Walls

Horizontal primary
control surface

dd

i

Bldg.
elevation
datum

10 ft 0 in. (3m)

Vertical primary
control surface

h10

a

a1

CL of steel
suport

Plan View
Spandrels

10 ft 0 in. (3m)

Vertical primary
control surface
e

e
b

k
f

c
Support
member
Side View
Spandrels

Bldg.
elevation
datum

g
Elevation View
Spandrels

The primary control surfaces are usually as shown, although this needs to be confirmed on a job-byjob basis.
a

=	Plan location from building grid datum* ................................................... ± 1/2 in. [± 13 mm]

a1 =	Plan location from centerline of steel support† ......................................... ± 1/2 in. [± 13 mm]
b

=	Top elevation from nominal top elevation:
Exposed individual panel ............................................................................. ± 1/4 in. [± 6 mm]
Nonexposed individual panel . .................................................................... ± 1/2 in. [± 13 mm]

c

=	Support haunch elevation from nominal elevation:
Maximum low . ................................................................................................... 1/4 in. [6 mm]
Maximum high . ................................................................................................ 1/2 in. [13 mm]

d

=	Maximum plumb variation over height of structure
or 100 ft [30 m] whichever is less* ..................................................................... 1 in. [25 mm]

e

= Plumb in any 10 ft [3 m] of element height ........................................................ 1/4 in. [6 mm]

f

=	Maximum jog in alignment of matching edges:
Exposed relative to adjacent panel ..................................................................... 1/4 in. [6 mm]
Nonexposed relative to adjacent panel . ........................................................... 1/2 in. [13 mm]

g

= Joint width (governs over joint taper) .......................................................... ± 1/4 in. [± 6 mm]

h

= Joint taper maximum ........................................................................................ 3/8 in. [10 mm]

h10 =	Joint taper over 10 ft [3 m] length ...................................................................... 1/4 in. [6 mm]
i

=	Maximum jog in alignment of matching faces . .................................................. 1/4 in. [6 mm]

j

=	Differential bowing or camber as erected between
adjacent members of the same design . .............................................................. 1/4 in. [6 mm]

k

=	Opening height between spandrels ............................................................. ± 1/4 in. [± 6 mm]

*For precast concrete buldings in excess of 100 ft [30 m] tall, tolerances “a” and “d” can increase at
the rate of 1/8 in. [3 mm] per story to a maximum of 2 in. [50 mm].
†For precast concrete elements erected on a steel frame, this tolerance takes precedence over tolerances
on dimension “a”.

360

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.6.3 Erection Tolerances

4

Fig. 4.6.11 Column erection tolerances.

d d

a

Bldg. grid
datum

a

a
a

Plan view
Splice
Vertical primary
control surface
100' –0" (30m) max.

f

e

e

10' (3m)

Horizontal primary
control surface
(at first corbel)

b
Bldg. elevation
datum

c

Bldg. elevation
datum
Elevation

a

Bldg. grid
datum
Elevation

The primary control surfaces are usually as shown, although this needs to be confiremed on a job-byjob basis.
a

=	Plan location from building grid datum:
Structural applications ................................................................................... ± 1/2 in. [13 mm]
Architectural applications ................................................................................ ± 3/8 in. [9 mm]

b

=	Top elevation from nominal top elevation:
Maximum low . ................................................................................................. 1/2 in. [13 mm]
Maximum high . .................................................................................................. 1/4 in. [6 mm]

c

=	Bearing haunch elevation from nominal elevation:
Maximum low . ................................................................................................. 1/2 in. [13 mm]
Maximum high . .................................................................................................. 1/4 in. [6 mm]

d

=	Maximum plumb variation over height of element (element in
structure of maximum height of 100 ft [30 m]) . ................................................. 1 in. [25 mm]

e

= Plumb in any 10 ft [3 m] of element height ........................................................ 1/4 in. [6 mm]

f

=	Maximum jog in alignment of matching edges:
Architectural exposed edges ............................................................................... 1/4 in. [6 mm]
Visually non-critical edges ................................................................................ 1/2 in. [13 mm]

ARCHITECTURAL PRECAST CONCRETE

|

361

4

DESIGN

4.6.3 Erection Tolerances

If reasonable tolerances and adjustments have been
designed into the construction details, more precise installation and general improvement in appearance are
achieved, and the erector should be able to:
1. Avoid joint irregularities, such as tapered joints
(panel edges not parallel), jogs at intersections,
and non-uniform joint widths.
2. Maintain proper opening dimensions.
3. Properly execute all fastening connections.
4. Align the vertical faces of the units to avoid outof-plane offsets.
5. Adjust for the accumulation of tolerances.
The precast concrete erector should perform a survey
of the building as constructed and lay out joint centerlines spaced along an elevation prior to actual product
installations and center the units between them. This
will keep the differential variation in joint width to a
minimum, as well as identifying problems caused by
building frame columns or beams being out of dimension or alignment. Horizontal and vertical joints should
be aligned and uniform joint widths should be maintained as erection progresses.
Variations from true length or width dimensions of
the overall structure are normally accommodated in
the joints or, where this is not feasible or desirable,
at the corner panels, in expansion joints, or in joints
adjacent to other wall materials. A liberal joint width
should be allowed if variations in overall building dimensions are to be absorbed in the joints. This may be
coupled with a closer tolerance for variations from one
joint to the next for uniformity of appearance purposes. The individual joint width tolerance should relate to
the number of joints over a given building dimension.
For example, to accommodate reasonable variations
in actual site dimensions, a 3/4 in. (19 mm) joint may
be specified with a tolerance of ± 1/4 in. (±6 mm) but
with only a 3/16 in. (5 mm) differential variation allowed
between joint widths on any one floor or between adjacent floors.
In a situation where a joint must match an architectural feature (such as a false joint), a large variation
from the theoretical joint width may not be acceptable
and tolerance for building lengths may need to be accommodated at the corner panels. A similar condition
often occurs where precast concrete is interspersed
with glass or curtain wall elements, as in precast concrete mullion projects. Close tolerances are often man-

362

|

ARCHITECTURAL PRECAST CONCRETE

datory between the mullion and the glass or curtain
wall. This condition demands additional flexibility that
may be provided by the corner details.
Clearance is the space provided between the structure and precast concrete members. It is one of the
most important factors to consider in erection because
of its impact on the final appearance of the structure.
The clearance space should provide a buffer area where
frame, erection, and product tolerance variations can
be absorbed. Clearances should be reviewed during
the design stages of the project to ensure they are appropriate from both erection and aesthetic points of
view.
With reasonable tolerances for the building frame
established, it is equally important that the designer
provide adequate clearances, for example, between
the theoretical face of the structure and the back face
of a precast concrete panel in detailing the panel and
its relationship to the building structure. If clearances
are realistically assessed, they will enable the erector
to complete the final assembly without field-altering the physical dimensions of the precast concrete
units. Adjacent materials may include products such
as glass or subframes that are installed after the precast concrete panels are in place. If sufficient space is
not provided, alignment of the wall as specified will
likely necessitate delays and extra costs and may be
impossible.
The failure to provide adequate clearances may cause
problems during wall installation. When determining
clearances, the following primary basic considerations
should be addressed:
• Product tolerance

• Erection tolerance

• Type of member

• Fireproofing of steel

• Size of member

• Thickness of plates,
bolt heads, and other
projecting elements

• Location of member
• Member movement
• Function of member

• Working space to make
the connection

Designing clearances should consider not only the
dimensional tolerance of the precast concrete members, but also the dimensional accuracy of the support
system (building frame). Clearances must enable the
erector to complete the final assembly without field altering the physical dimensions of the precast concrete
units.

DESIGN

4.6.3 Erection Tolerances / 4.6.4 Interfacing Tolerances

The type of member is partially accounted for when
product tolerances are considered. There are additional
factors that should also be considered. An exposedto-view member requiring small erection tolerances
requires more clearance for adjustment than a non-exposed member with a more liberal erection tolerance.
Similarly, a corner member should have a large enough
clearance so it can be adjusted to line up with both of
the adjacent panels.
The size and weight of the member are other considerations in determining erection clearances. Large
members are more difficult to handle than smaller
ones; a large member being erected by a crane requires more clearance than small member that can be
hand-erected or adjusted.
Clearance design should consider member deflection, rotation, and movements caused by temperature expansion and contraction, creep, and shrinkage.
Clearance between a vertical member and a horizontal
member should allow for some movement in the horizontal member to prevent the vertical member from
being pushed or pulled out of its original alignment.
If not considered in the design, such movements can
create waterproofing problems or roofing failure at the
interface.
Consideration should be given to the limits of adjustment permitted by the connection details. All connections should provide maximum adjustability in all directions that is structurally or architecturally feasible.
When a 1 in. (25 mm) clearance is needed but a 2
in. (50 mm) clearance creates no structural or architectural difficulty, the 2 in. (50 mm) clearance should
be selected. Closer tolerances are required for bolted
connections than for most grouted connections. To accommodate any misalignment of the building frame,
connections should provide for vertical, horizontal, and
lateral adjustments of at least 11/2 in. (38 mm). If a connection is attached to a spandrel beam or column that
is fireproofed, more clearance will be needed to install
fastenings than when the anchors are located on the
top surface of beams and the sides of columns. Also,
space should be provided to make the connection (sufficient room for welding or adequate space to place a
wrench to tighten a bolt).
Nominal clearance dimensions shown on the erection drawings should be equal to the actual clearance
required plus the outward tolerance permitted for
the adjacent construction, and should be determined

4

based on the assumption that the construction will be
as far out of position in the wrong direction as is allowed. Special attention should be given to complex
geometric interfaces. Connections should be designed
to accommodate the clearance plus tolerance.
If the clearance provided is too tight, erection may
be slow and costly because of fit-up problems and the
possibility of rework. A good rule of thumb is that at
least 3/4 in. (19 mm) clearance should be required between precast concrete members, with 1 in. (25 mm)
preferred; 11/2 in. (38 mm) is the minimum clearance
between precast concrete members and cast-in-place
concrete, with 2 in. (50 mm) preferred. For steel structures, 11/2 in. (38 mm) is the minimum clearance between the back of the precast concrete member and
the surface of the fireproofing, with 2 in. preferred.
If there is no fireproofing required on the steel, then
11/2 in. (38 mm) minimum clearance should be maintained. At least a 11/2 to 2 in. (38 to 50 mm) clearance
should be specified in tall structures, regardless of the
structural framing materials. The minimum clearance
between column covers and columns should be 11/2
in. (38 mm); 3 in. (75 mm) is preferred because of the
possibility of columns being out of plumb or a column
dimension causing interference with completion of the
connection. If clearances are realistically assessed, they
will solve many installation problems. Where large tolerances have been allowed for the supporting structure, or where no tolerances for the structure are given, the clearance may have to be increased.

4.6.4 Interfacing Tolerances
Interfacing tolerances and clearances are required for
the joining of different building materials in contact
with or in close proximity to precast concrete, and to
accommodate the relative movements expected between such materials during the life of the building.
Typical examples include tolerances for window and
door openings; joints, flashing, and reglets; mechanical and electrical equipment; elevators and interior finishes; and walls and partitions.
Fabrication and erection tolerances of other building
materials must also be considered in design of the precast concrete units and coordinated to accommodate
the other functional elements comprising the total
structure. Unusual requirements or allowances for interfacing should be stated in the contract documents.
When the matching of different building elements

ARCHITECTURAL PRECAST CONCRETE

|

363

4

DESIGN

4.6.4 Interfacing Tolerances / 4.7 Joints

is dependent on work executed at the construction
site, interface tolerances should be related to erection tolerances. When the execution is independent of
site work, tolerances should closely match the normal
manufacturing tolerances for the materials to be joined
plus an appropriate allowance (clearance) for differential volume changes between materials. For example,
window elements have installation details that require
certain tolerances on window openings in a precast
concrete panel. If the opening is completely contained
within one panel, can the required tolerances on the
window opening be economically met? If not, is it
less expensive to procure special windows or to incur
the added cost associated with making the tolerances
on the window opening more stringent? Also, openings for aluminum windows should allow clearance for
some thermal expansion of the frame.
It is important to note that interfacing tolerances may
be system dependent. For example, windows of one
type may have a different interface tolerance than windows of another type. If material or component substitutions are made after the initial design is complete,
the responsibility for ensuring that the interface tolerances are compatible with adjacent building materials
passes to the party initiating the substitutions.
Adequate interface/erection tolerances are required
for window openings, doors, or louvers common to
two or more panels. The cost of erecting the panels
to achieve required window interface tolerances must
also be considered. A similar condition often occurs
where precast concrete is interspersed with glass or
metal curtain wall elements, as in many precast concrete spandrel or mullion projects. Close tolerances are
often mandatory between the mullion and the glass or
curtain wall. This condition demands additional flexibility that may be provided in the corner details. Also,
any bow in spandrel panels is critical if windows are to
be installed between panels.
Product tolerances, erection tolerances, and interface tolerances together determine the dimensions of
the completed structure. Which tolerance takes precedence is a question of economics, which should be
addressed by considering fabrication, erection, and interfacing cost implications.
Special tolerances or construction procedures require
early decisions based on overall project economics.
Once these decisions have been made, they should
be reflected in the project plans and specifications. All

364

|

ARCHITECTURAL PRECAST CONCRETE

special tolerance requirements or allowances for interface, special details, and special procedures should be
clearly spelled out in the specifications. The plans and
specifications then define the established tolerance
priority for the project.

4.7 JOINTS
4.7.1 General
The successful performance of a building exterior is
frequently defined by its ability to keep rain and the
elements outside, away from the building’s occupants.
Precast concrete panels are relatively impermeable to
water. Moisture will not penetrate through precast
concrete panels. The joints between precast concrete
panels or between panels and other building materials
must be considered to prevent water and air penetration through the building envelope. The design and
execution of these joints is therefore of the utmost
importance and must be accomplished in a rational,
economical manner. Joint treatment also has an effect
on the general appearance of the project. To ensure
the joint and sealant give the desired performance, selecting the right product, appropriate joint design, and
proper surface preparation and application technique
is required.
The penetration of moisture into a building envelope
may enter directly (through an opening), by gravity,
capillary action, and as a result of the mean (steady
state) air pressure difference across the wall.
Joint sealants are fully exposed to the major agents of
aging and deterioration—ultraviolet light and thermal
cycling. High-performance sealants with a low modulus and high movement capability must be used to ensure quality long-term performance. In new construction, labor to material costs are typically 4 to 1, while
in renovation/rehabilitation the ratio may be 8 to 1 or
more.
Joints are required to accommodate changes in wall
panel or structure dimensions caused by changes in
temperature, moisture content, or deflection from
applied design loads. The joints between panels are
normally designed to accommodate local wall movements rather than cumulative movements. Sealants
subjected to volume change movements, either horizontally or vertically at building corners, at adjacent
non-precast concrete construction, or at windows not
having similar movements must be given special con-

DESIGN

4.7 Joints / 4.7.2 Types of Joints

sideration. Some wall designs handle water properly
in two-dimensional blueprints, but fail in three-dimensional reality. Isometric drawings should be used to
show the proper intersection of horizontal and vertical
seals. These intersections are a prime source of sealant
problems.

Fig. 4.7.1 Single stage joint.
A
Tooled surface

B

4.7.2 T ypes of Joints

C

Joints between precast concrete wall units may be divided into three basic types: one-stage, two-stage, and
expansion joints.
One-Stage joint–As its name implies, the onestage (face-sealed) joint has a single line of caulking
for weatherproofing. This is normally in the form of a
gun-applied sealant close to the exterior surface of the
precast concrete panel (Fig. 4.7.1).
The principal advantages of face-sealed joints are
their simplicity, ease of installation, and almost universal suitability for normal joints between precast
concrete panels. No grooves or special shapes are necessary. Thus, one-stage joints are normally the most
economical with regard to initial cost. However, the
economics may change when maintenance costs are
included in the evaluation. One-stage joints provide
adequate air leakage and water penetration control in
most climates. Their performance depends greatly on
the quality of sealant materials, the condition of joint
surfaces, quality of field installation, and the overall
wall design.
Because sealants are subject to deterioration from
the elements and ultraviolet (UV) exposure, it is recommended that the sealant be set back into the joints by
using recessed joints. This partially protects the sealant
from rain, wind, and UV light.
Two-Stage joint–Watertightness of sealant joints
can be improved by installing a second line of sealant
in each joint. The inner seal is placed inside the joint,
generally from the exterior, and recessed a minimum
of 2 to 21/2 in. (50 to 63 mm) from where the back of
the front sealant and backing will be located or to the
back of insulation in a sandwich (insulated) wall panel.
This layer provides redundancy in the system, as it is
fully protected from weather and UV exposure by the
outer layer of sealant, which is installed in the normal
manner.
This approach requires the installation of 3/8 in. (10
mm) weep openings in the exterior seal to allow water

4

Key Points:
1. Dimension C must be at least 1/4 in.
2. Ratio of A:B should be 2:1 minimum
3. Joint surface tooled concave.
4. Dimension B = 3/8 in. with a minimum of 1/8 in.
over the crown of backer rod.
5. Dimension A = 3/4 in. minimum recommended.

Sealant

Pre-formed
rod or tubular
backing. If this
material is nonadhesive, bondbreaker is not
necessary.

Note: 1 in. = 25.4 mm

contained by the inner seal to exit the cavity between
joint seals. Near the junction of the horizontal and vertical joints, the inner seal must turn out to the plane
of the exterior seal at regular intervals to force water
out of the joint (Fig. 4.7.2). This termination requires
care in detailing and construction. Failure to provide
these weep openings results in trapped water within
the joint and ponding against both seals; this accelerates deterioration of the sealant material and its bond
to the substrate.
These joints are based on the open rainscreen principle. They are sometimes known as ventilated or pressure equalization joints and are favored for exterior
wall construction in Canada. The rainscreen principle
is based on the control of the forces that can move
water through small openings in a face-sealed wall
system, rather than the elimination of the openings
themselves.
These joints have two lines of defense for weatherproofing. The typical joint consists of a rain barrier near
the exterior face and an air retarder close to the interior
face of the panel. The rain barrier is designed to shed
most of the water from the joint, and the wind-barrier
or air retarder is the demarcation line between outside
and inside air pressure.
Openings in the rain barrier allow air to rapidly enter

ARCHITECTURAL PRECAST CONCRETE

|

365

4

DESIGN

4.7.2 Types of Joints / 4.7.3 Expansion Joints

until the pressure inside the chamber is equal to the
wind pressure acting against the outer wall, which prevents water from entering the chamber. The pressure
difference across the exterior layer is essentially zero,
and wind pressure is transferred to the inner, airtight
layer. Rain does not penetrate to the air chamber and,
subsequently, to the interior of the building because
there is no wind pressure forcing it through the exterior layer. Any moisture entering the joint will cling to
the joint walls and then be drained out by the transverse seal.
The airtightness of the air retarder is critical in governing the speed at which pressure equalization occurs. Pressure equalization must take place almost instantaneously for a rainscreen wall to be effective. The
size of vent opening must reflect the size of the joint to
be pressure equalized.
Typical details of two-stage vertical joints are shown
in Figs. 4.7.2 and 4.7.3. This system is especially applicable to high-rise buildings subject to severe climatic exposure (greater than 5000 degree days). The
warm, moist air moving from the building interior to
the exterior usually carries moisture, which could cause
condensation. Air must be prevented from contacting
cold surfaces in the wall. In northern climates, thermal bridges can occur and allow condensation to form
a buildup of frost in or on the walls, which may be
thought to be a failure of the joint sealant. This frost
later can melt and run back inside the building, giving
the impression that the building is leaking.
Water either from penetration or condensation in
the joint should be drained from the joint by proper
sealant installations. The second line of sealant should
be brought to the front face at regularly spaced intervals along the height of vertical joints, usually near the
junction of the horizontal and vertical joints at each
floor level. Therefore, if any moisture does come out of
the system, it will run down the face of the joint sealant and not over the face of the panels.
A spacing of two or three stories may be sufficient
for low-rise buildings and in areas of moderate wind
velocities. Factors to consider when using two-stage
joints are:
1. H
 igher initial cost due to labor and materials required for their successful application.
2. S ealants are not easily placed at the back of the
two-stage joint unless 1 to 13/8 in. (25 to 35 mm)

366

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 4.7.2 Two-stage vertical joints.

Backer Rod

Inner Seal
Outer Seal

Weep Opening To Drain Cavity
Between Seals
Backer
Rod
Caulking

Drain/Vent

joints are used. Therefore, conscientious workers
or intensive supervision throughout the installation
procedure is necessary, because inspection of the
completed installation is difficult.
Panel configurations and joint widths should permit
a careful applicator to successfully install both lines of
sealant from the exterior. The special tools required
may include an extension for the nozzle of the caulking
gun and a longer tool for tooling the interior sealant.
The architect, precaster, erector, and sealant applicator must all understand the function of the two-stage
joints if optimum results are to be achieved. The dimensions of the joints must be maintained at all times.
The most common mistake in the installation of twostage joints is leaving gaps in the air seal.

4.7.3 Expansion Joints
Cumulative movements, as well as differential expansion movement of adjacent wall materials, are generally
taken by specially designed expansion joints. Because an

DESIGN

4.7.3 Expansion Joints

4

Fig. 4.7.3 Sealant and joint details.
Room Finish
Air / Vapor
Barrier
Insulation

Room Finish
Air / Vapor
Barrier
Insulation

Interior Air Barrier Sealant
Bead Installed From Exterior

Interior Air
Barrier Sealant
Bead Installed
From Exterior

Both Vertical &
Horizontal Sealant
Beads Interface For
Complete Air Seal

Exterior
Weather
Seal

/ in. Minimum

3 4

Vent / Weep
Hole

Transverse
Drainage Bead

Both Vertical
& Horizontal
Interface For
Complete Air
Seal
Plan Section At Horizontal Joint

Transverse
Drainage Bead

Vent Weep Hole

Exterior
Weather Seal

Vertical Section At Window

Vertical Section At Vertical Joint
Insulation
Insulation

Structural
Wythe
Exterior
Wythe

Exterior
Wythe
Interior
Air / Vapor
Sealant Bead

Optional Sealant
Bead For Interior
Aesthetics (Vented
To Room)

Vent / Weep
Hole

1 in. Minimum
Exterior
Weather
Seal

Transverse
Drainage Bead
Optional
Room Finish
Exterior Weather
Seal

Plan Section At Horizontal Joint

expansion joint may have to accommodate considerable
movement, it should be designed as simply as possible.
Although this might result in an appearance somewhat
different from a normal joint, the architect is urged to
either treat it as an architectural feature or simply leave
it as a different, but honest, expansion joint.

Vertical Section At Vertical Joint

Seismic seals are a special case of expansion joints.
Such joints are generally quite large and are used between new and existing buildings to protect the joint
from moisture and allow the structures to move from
thermal expansion, wind drift, and seismic motions
without damage. Seismic joints are designed to ac-

ARCHITECTURAL PRECAST CONCRETE

|

367

4

DESIGN

4.7.3 Expansion Joints / 4.7.5 Location of Joints

economy by working with larger panels.
Fig. 4.7.4 Expansion seal in vertical joint system.
Inner Seal Closed
Cell Neoprene
3/16” Dia. Exp.
Bolt @ 24” O.C.
(5 mm Dia. @
610 mm O.C.)

Adjacent Construction

Extruded Alum.
Retainer

Butyl Sealant

Nominal Jt. Width
Outer Seal
Dense Neoprene

Tear Strip
3/16” Dia. Exp. Bolt @ 24” O.C.
(5 mm Dia. @ 610 mm O.C.)
Inner Seal Closed
Cell Neoprene

Extruded Alum.
Retainer
Butyl Sealant

Outer Seal
Dense Neoprene

Adjacent
Construction
Nom. Jt. Width
Tear Strip

Pleat

Note: Location and number of tear strips vary with joint size;
1 in. = 25.4 mm.

commodate both vertical and horizontal movement. They
are available in sizes from 2 to 12 in. (50 to 305 mm).
Wider openings can be accommodated by joining seal sizes
together.
Materials for expansion joints must be chosen for their
ability to absorb appreciable movement while performing
their primary function of controlling the movement of moisture and air. Figure 4.7.4 shows bellows-type expansion
seals of neoprene that accommodate thermal movement
and seismic movement. Joints must be designed first for
weather protection longevity, then for movement, and finally for appearance. In most cases, this requires that special
gasket materials be used, rather than sealants. Otherwise,
the requirements for expansion joints are similar to those
listed previously for other joints.

4.7.4 Number of Joints
The number of joints in the architectural design should
be minimized. This will result in a lower overall-cost for the
joints, potentially lower maintenance costs, and will increase

368

|

ARCHITECTURAL PRECAST CONCRETE

Limiting panel sizes to minimize movements in the joints is
not recommended. It is generally more economical to select
larger panels and design the joints and sealants to allow
for anticipated movements. Optimum panel size should
be determined by erection conditions, available handling
equipment, and local transportation limitations as to panel
weight and sizes (see Section 3.3.9).
If the desired appearance demands additional joints, false
joints may be used to achieve a more balanced architectural appearance. In order to match appearance of the two
joints, the finish of the false joints should simulate the gaskets or sealants used in the real joints. Caulking false joints
adds unnecessary expense.

4.7.5 L ocation of Joints
Joints are simpler to design and execute if they are
located at the maximum panel thickness. If there are
any ribbed projections at the edges of the panels, joints
should be placed at this location. Ribs at the edges
improve the structural behavior of the individual unit.
Also, panel variations—possible warping or bowing—
are less noticeable when the joints are placed between
ribs than when the joints are located in flat areas.
However, complete peripheral ribs are not recommended because they are likely to cause localized water runoff resulting in unsightly staining. Instead, ribs should
be placed at vertical panel edges. If the ribs are too narrow to accommodate joints, the full rib may be located
in one panel only.
Vertical joints should be located on grid lines. Horizontal
joints should be near, but above, floor lines. The designer
should allow the precaster to optimize panel sizes for economy with false joints, if necessary. The location of joints between precast concrete panels should be considered as an
integral part of the evaluation of economical fastening of
the units.
Locating and detailing joints (real or false) is an important
factor in creating weathering patterns for a building. Joints
should be made wide and recessed to limit unexpected
weathering effects (Fig. 4.7.5). Recessed joints screen the
joint from rain by providing a dead-air space that reduces
air pressure at the face of the sealant. Also, the joint profile channels the rain runoff, helping to keep the building
façade clean from unsightly runoff patterns. The designer
should determine where the water will finally emerge. Setbacks should be provided at window perimeters and other

DESIGN

4.7.5 Location of Joints / 4.7.6 Width and Depth of Joints

Fig. 4.7.5 Typical architectural panel joints.

4

Fig. 4.7.6 Proper channeling of water.

DON’T

Chamfered

Quirk Miter

Joint

Recessed

DO
False Joint
Alternate Recessed

Return With Real And False Joint

Fig. 4.7.7 Staggered architectural wall panels.

vulnerable joints in the wall system to reduce the magnitude
and frequency of water exposure.
Figure 4.7.6 shows an elevation where some of the false
vertical joints, into which water is channeled, discharge this
water over a vertical concrete surface with fewer joints than
at higher levels. This causes a marked washing effect at termination of the joint; the water should be directed until it
reaches the ground or a drainage system.
Joints in forward-sloping surfaces are difficult to weatherproof, especially if they collect snow or ice. This type of
joint should be avoided, whenever possible. When forward
sloping joints are used, the architect should take special precautions against water penetration.
All joints should be aligned, rather than staggered,
throughout their length (Fig. 4.7.7). Non-aligned joints
subject sealants to shear forces in addition to the expected
compression or elongation forces. The additional stress may
cause sealants to fail. In addition, non-aligned joints force
panels to move laterally relative to each other, inducing high
tensile forces.

4.7.6 Width and Depth of Joints
Joint width must not only accommodate variations in the
panel dimensions and the erection tolerances for the panel,
but must also provide a good visual line and sufficient width
to allow for effective sealing.
The performance characteristics of the joint sealant should
be taken into account when selecting a joint size. Joints be-

Shadow Lines

Staggered Architectural Panels

tween precast concrete units must be wide enough to accommodate anticipated thermal expansion, as well as other
building movements and proper sealant installation. Joint
tolerances must be carefully evaluated and controlled if the
joint sealant system is to perform within its design capabilities. When joints are too narrow, bond or tensile failure of
the joint sealant may occur and/or adjacent units may come
in contact and be subjected to unanticipated loading, distortion, cracking, and local crushing (spalling).
Joint widths should not be chosen for reason of appearance alone, but must relate to panel size, building tolerances, joint sealant materials, and adjacent surfaces. The required width of the joint is determined by the temperature
extremes anticipated at the project location, the movement
capability of the sealant to be used, the temperature at
which the sealant is initially applied, panel size, fabrication
tolerances of the precast concrete units and panel installation methods. The following factors take precedence over
ARCHITECTURAL PRECAST CONCRETE

|

369

4

DESIGN

4.7.6 Width and Depth of Joints

appearance requirements:
1. Temperature extremes and gradients. The temperature range used when selecting a sealant must reflect
the differential between seasonal extremes of temperature and temperature at the time of sealant application. Concrete temperatures can and normally will vary
considerably from ambient air temperatures because of
thermal lag. Although affected by ambient air temperatures, anticipated joint movement must be determined
from anticipated concrete panel temperature extremes
rather than ambient air temperature extremes.
2. Sealant movement capability. A sealant’s performance within joints is rated as the allowable movement
expressed as a percentage of the effective joint width.
The minimum design width of a panel joint must take
into account the total anticipated expansion and contraction movement of the joint and the movement capability of the sealant. This evaluation should include
volume changes from creep, shrinkage, and temperature variations.
PCI Design Handbook supplies figures for estimating volume changes directly related to the size of the panel. Most
drying shrinkage occurs in the first weeks following casting,
and creep normally levels out after a period of months. For
these reasons, movements caused by ambient air temperature variations are more important than those caused by
shrinkage. For loadbearing panels, the effect of creep may
be cumulative, thus may be more important.
Many factors may be involved in actual building joint
movement. These include, but are not limited to, mass of
material, color, insulation, building load, building settlement,
method of fastening and location of fasteners, differential
heating due to variable shading, thermal conductivity, differential thermal stress (bowing), building sway, and seismic
effects. Material and construction tolerances that produce
smaller joints than anticipated are of particular concern.
Tolerances in overall building width or length are normally
accommodated in panel joints, making the overall building
size tolerance an important joint consideration. Where a
joint must match an architectural feature (such as a false
joint), a large variation from the theoretical joint width may
not be acceptable and tolerances for building lengths may
need to be accommodated at the corner units.
A practical calculation of panel joint size can be made as
follows, as shown in ASTM C1193 and C1472:
J=

370

|

100A
+B+C
x

ARCHITECTURAL PRECAST CONCRETE

where:
J

= minimum joint width, in.

X = stated movement capability of the sealant, in
percent
A = calculated movement of panel from thermal changes
= (coefficient of thermal expansion) (change in temperature) (panel length)
B = material construction tolerances
C = seismic or other considerations as appropriate
Example: Concrete panels of 30 ft (9.1 m) in length,
expecting a temperature change in the concrete of
60 °F (33 °C) from sealant installation temperature, with a
material or construction tolerance of 0.25 in. (6 mm), are
to be sealed with a sealant having ±50% movement capability (as determined by ASTM C719). The coefficient of
thermal expansion of the concrete is 6 × 10-6 in./in./°F. The
calculated movement of the panel from thermal change is
as follows:
A = (6 × l0-6 in./in./°F) (60 °F) (360 in.) = 0.130 in.
(3 mm)
X = 50%
B = 0.25 in. (6 mm)
No seismic considerations, (C = 0).
The calculated minimum joint width is as follows:
J=

(100)(0.130 in.)
+ 0.25 in. = 0.51 in. (13 mm)
50

To provide optimum quality for the installation and performance of sealants, the architect should specify a minimum
panel joint width of not less than 3/4 in. (19 mm). This is
the minimum nominal joint width needed to adequately account for production and erection tolerances and still maintain an effective minimum joint width that can be caulked.
The use of larger joints at reentrant corners and mitered
panels at outside corners helps to relieve the possibility of
impact between panels under large drifts in high seismic areas. It is also important that the joint between precast concrete panels and window frames also maintains the same
nominal joint width. Corner joints may be 11/4 in. (30 mm)
wide to accommodate the extra movement and bowing often experienced at this location. A minimum joint width of
3
/4 in. (19 mm) also is recommended for two stage joints to
allow sufficient space for insertion of the interior seal with
a 1 in. (25 mm) joint width recommended for insulated
panels.

DESIGN

4.7.6 Width and Depth of Joints / 4.7.7 Sealant Materials and Installation

The required sealant depth is dependent on the sealant
width at the time of application. The optimum sealant
width/depth relationships are best determined by the sealant manufacturer, however, generally accepted guidelines
are:
1. For joints designed for 3/4 to 1 in. (19 to 25 mm) width:
The sealant depth should be equal to one half the width.
The sealant should have a concave shape providing
greater thickness at the panel faces. The sealant should
have a minimum 1/4 in. (6 mm) contact with all bonding
surfaces to ensure adequate surface adhesion.
2. For joints greater than 1 in. (25 mm) wide: Sealant
depth should be limited to 1/2 in. (13 mm) maximum,
preferably 3/8 in. (10 mm). For sealant widths exceeding 2 in. (50 mm), the depth should be determined by
consultation with the sealant manufacturer.
The depth of the sealant should be controlled by using a
suitable sealant backing material. To obtain the full benefit
of a well-designed shape factor, the backing material must
also function as a bondbreaker (Fig. 4.7.1). When it comes
to sealant depth, more is not better. If too much sealant is
applied, the stresses on the sealant bead are magnified and
the chance of premature debonding at the precast concrete
interface is increased. If the bead is too shallow, there may
be insufficient material to accommodate the joint movement and the sealant will split.

4

rather than generic sealant type.
When specifying a sealant, a current sample warranty
should be obtained from the manufacturer and the contents
studied to avoid uncalculated risks. The warranty period for
a polyurethane material can be up to 10 years, and up to
20 years for a silicone. This doesn’t imply that the sealant
will deteriorate during that time. Some polyurethane-based
products maintain their appearance and integrity for more
than 15 years. Warranties can be written to cover either the
material or the material and the labor needed to replace
them. The specifier should be familiar with the available
sealants and associated warranties prior to selecting a sealant for the building.
The following characteristics should be considered when
making the final selection of sealants from those with suitable physical (durability) and mechanical (movement capability) properties:
1. Adhesion to different surfaces—concrete, glass, or
aluminum.
2. Surface preparation necessary to ensure satisfactory
performance—priming, cleaning, and drying.
3. Serviceable temperature range.
4. Drying characteristics—dirt accumulation, susceptibility to damage due to movement of joint while sealant
is curing.
5. Puncture, tear, and abrasion resistance.

4.7.7 Sealant Materials and Installation

6. Color and color retention.

The most common joint materials are sealants meeting
ASTM C920. These sealants are used in both one-stage and
two-stage joints. If used as an air seal, they may be applied
from the front provided joint width and depth permit, or
from the interior if access to the joint is not blocked by edge
beams or columns.

8. Staining of adjacent surfaces caused by sealant or
primer.

Designers should consult with the various sealant suppliers
to ensure they are specifying an appropriate sealant for the
specific needs of the project, as well as the sealant’s proper
installation. For a comprehensive discussion of joint sealants
used between wall panels, refer to ASTM C1193, Standard
Guide for Use of Building Sealants. Table 4.7.1 provides a
list of common sealants and their qualities. Non-staining
joint sealants should be selected to prevent the possibility of
bleeding and heavy dirt accumulation, which are common
problems with sealants having high plasticizer contents.
Also, care should be taken to avoid sealants that collect dirt
as a result of very slow cure or long tack-free time. Dirt accumulation is more a function of specific product formulation

7. Effect of weathering—water and ultraviolet (UV) light—
on properties such as adhesion, cohesion, elasticity.

9. Ease of application.
10. Environment in which the sealant is applied.
11. Compatibility with other sealants to be used on the
job.
12. Long term durability.
13. Life expectancy.
The sealants used for specific purposes are often installed
by different subcontractors. For example, the window
subcontractor normally installs sealants around windows,
whereas a different subcontractor typically installs sealants
between panels. The designer must select and coordinate
all of the sealants used on a project for chemical compatibility and adhesion to each other. In general, contact between different sealant types should be avoided by having

ARCHITECTURAL PRECAST CONCRETE

|

371

4

DESIGN

4.7.7 Sealant Materials and Installation

Table 4.7.1. Comparative Characteristics and Properties of Field-Molded Sealants.

Polysulfides
OneTwoComponent
Component
Base:
Polysulfide
polysulfide
polymers,
polymers,
activators,
activators,
pigments,
pigments,
Chief ingredients
inert fillers,
plasticizers,
curing agents, fillers Activator:
nonvolatilizing accelerators,
extenders,
plasticizers
activators
Primer required
Usually
Usually
Chemical
Chemical
reaction with
reaction with
Curing process
moisture in air
curing agent
and oxidation
Tack-Free time, hr (ASTM C679)
24
36 – 48
Cure time, days1
7 – 14
7
Max. cured elongation (ASTM
300%
600%
D412)
Recommended max. joint
± 25%
± 25%
movement (ASTM C719)
3
Max. joint width, in.
/4
1
2
Resistance to compression
Moderate
Moderate
Resistance to extension2
Moderate
Moderate
Service temperature range, °F
– 40 to + 200
– 60 to + 200
Normal application temperature
+ 40 to + 120
+ 40 to + 120
range, °F
Weather resistance
Good
Good
Ultra-Violet resistance, direct
Good
Good

Polyurethanes
OneTwoComponent
Component

Silicones
OneComponent

TwoComponent

Base:
polyurethane
prepolymer,
filler pigments,
plasticizers
Activator:
accelerators,
extenders,
activators

Siloxane polymer
pigments:
alcohol or other
non-acid cure

Siloxane
polymer
pigments:
alcohol or
other non-acid
cure

Usually

Usually

Occasionally

Occasionally

Chemical
reaction with
moisture in air

Chemical
reaction with
curing agent

Chemical
reaction with
moisture in air

Chemical
reaction with
curing agent

24 – 36
7 – 14

24 – 72
3–5

1–2
7 – 14

300%

500%

400 – 1600%

400 – 2000%

± 15%

± 25%

1¼
High
Medium
– 40 to + 180

2
High
Medium
– 40 to + 180

± 25% to
+ 100, - 50%
3
Low
Low
– 60 to + 250

± 121/2%
to ± 50%
3
Low
Low
– 60 to + 250

+ 40 to + 120

+ 40 to + 120

– 20 to + 110

– 20 to + 110

Very good
Poor

Very good
Poor

Excellent
Excellent

Polyurethane
prepolymer,
filler pigments,
plasticizers

/2 – 5
4–7

1

Excellent
Excellent
Excellent
Cut, tear, abrasion resistance
Good
Good
Excellent
Excellent
Good – excellent
– knotty tear
20
20
10 – 20
10 – 20
20
20
Life expectancy, years3
Hardness, Shore A (ASTM C661)
25 – 35
25 – 45
25 – 45
25 – 45
15 – 35
15 – 40
FS: TT-S-00230C FS: TT-S-00227E FS: TT-S-00230C FS: TT-S-00227E FS: TT-S-00230C FS: TT-S-00227E
ASTM C920
ASTM C290
ASTM C920 FS-TT-S-001543A USASI A-116.1
ASTM C920
Applicable specifications
(19-GP-13A)
(19-GP-24)
(19-GP-24)
ASTM C920
(19-GP-13)
ASTM C920
(Canadian)
(19-GP-3B)
(19-GP-18)
(19-GP-19)

1

 ure time, as well as pot life, are greatly affected by temperature and humidity. Low temperatures and low humidity create longer pot life and cure time; conversely, high temC
peratures and low humidity create shorter pot life and cure time. Typical examples of variations are:

Two-Part Polysulfide
Air temperature, °F
Pot life, hours Initial cure, hours Final cure, days
50
7–14
72
14
77
3–6
36
7
100
1–3
24
5
2
Resistance to extension and compression is better known in technical terms as modulus, the unit stress required to produce a given strain. It is not constant but changes in
values as the amount of elongation changes.
3

L ife expectancy is directly related to joint design, workmanship, and conditions imposed on any sealant. The length of time illustrated is based on joint design within the limitations outlined by the manufacturer, and good workmanship based on accepted field practices and average job conditions. A violation of any one of the above would shorten
the life expectancy to a degree. A total disregard for all would render any sealant useless within a very short period of time. Note: °F = °C (1.8) + 32; 1 in. = 25.4 mm.

372

|

ARCHITECTURAL PRECAST CONCRETE

DESIGN

4.7.7 Sealant Materials and Installation

one sealant contractor do both panel and window sealant
application with compatible materials.
The recommendations of the sealant manufacturer should
always be followed regarding mixing, surface preparation,
priming, application life, and application procedure. Good
workmanship by qualified sealant applicators is the most
important factor required for satisfactory performance.
Sealant installation should be specified to meet at least the
requirements of ASTM C1193.
Prior to sealant application, the edges of the precast
concrete units and the adjacent materials must be sound,
smooth, clean, and dry. They must also be free of frost, dust,
laitance, or other contaminants that may affect adhesion,
such as form release agents, retarders, or sealers. It may be
more economical and effective to prepare joint surfaces prior
to erection if a large number of units require surface preparation. It may also be desirable to conduct pre-project adhesion tests in accordance with ASTM C794, “Test Method for
Adhesion-in-Peel of Elastomeric Joint Sealants,” and field
adhesion tests using ASTM C1521, “Standard Practice for
Evaluating Adhesion of Installed Weatherproofing Sealant
Joints,” to determine the adhesion of the sealant with each
contact surface. Adhesion (ASTM C794 or C1521) and stain
testing (ASTM C510 or C1248) of the substrates and sealants in the early project planning stage of a building are
recommended by most sealant manufacturers. This early
testing will prevent most problems before they start and will
give the construction team the assurance of a problem-free
job.
Even when performed on a limited basis, inspecting sealants during installation significantly improves the probability they will be installed in accordance with the contract
documents. Performing this evaluation early in the project
provides a method for obtaining feedback on installation
workmanship. This way, modifications or corrections can be
implemented before any problem becomes widespread.
ASTM C1521 provides guidance for two tests. The first
is non-destructive, and consists of applying pressure to the
surface of the sealant at the center of the joint and the
bond line with a probing tool. The second procedure involves removing sealant to evaluate adhesion and cohesion.
The latter test offers tail and/or flap procedures, depending
on whether similar or different substrates are present on adjacent surfaces of the sealant joint. The sealant pulled from
the test area should be repaired by applying new sealant to
the test area. Assuming good adhesion was obtained, use
the same application procedure to repair the areas as was
used to originally seal them. Care should be taken to ensure

4

that the new sealant is in contact with the original sealant
so that a good bond between the new and old sealants will
be obtained.
ASTM C1521 can be used to evaluate installed sealant
during mockups, at the start of work to confirm application methods, and throughout the work to confirm installation consistency. ASTM C1521 provides guidelines for the
frequency of destructive testing when evaluation is part of
a quality control program for a new installation. All results
should be recorded, logged, and sent to the sealant applicator and manufacturer for warranty issuance.
In the construction of a mockup for water penetration testing, the actual field construction techniques must be used.
If a leak develops, which usually occurs at the window to
precast concrete interface, the details need to be examined
and modified. Putting more sealant on to make the system
pass the test is not realistic, as this will generally not occur
during construction.
Sealants that chemically cure should not be applied to
wet or icy surfaces, as they may cure or set before they can
bond to the concrete surface. Some methyl methacrylate
resin sealers inadvertently sprayed in the joints may peel
away from the concrete surface, leaving a void between
sealant and concrete. Silicone water repellents in the joints
may prevent adhesion of sealants to the concrete surface.
Therefore sealant/sealer compatibility should be verified.
Abrasion cleaning using a stiff wire brush, light grinding, or
sandblasting followed by air blowing may be necessary to
remove surface contaminants. The sealant should be cured
14 days before applying water repellents. Care should be
taken to caulk first, as sealer may prevent proper adhesion
of sealant.
Also, before caulking, the joint may require solvent cleaning with a lint-free cloth dampened with an acceptable
cleaning-grade solvent followed by wiping with a dry cloth.
Isopropyl alcohol (IPA) is soluble in water and may be appropriate for winter cleaning, as it helps in removing condensation and frost by picking up surface moisture as it
evaporates. Xylene and toluene are not soluble in water and
may be better suited for warm weather cleaning. Follow
the solvent manufacturer’s safe handling recommendations
and local, state, and federal regulations regarding solvent
usage.
Sometimes, smooth concretes that are very shiny exhibit a
“skin” on the surface. The skin may peel off, leaving a gap
between it and the concrete after the joint sealant has been
applied to the concrete. It may be necessary to remove the

ARCHITECTURAL PRECAST CONCRETE

|

373

4

DESIGN

4.7.7 Sealant Materials and Installation

skin by using a stiff wire brush followed by a high-pressure
water rinse. The joint must be dry before applying the sealant. Wet concrete should be allowed to dry for at least 24
hours, under good drying conditions, before applying sealant or primer.
The caulking gun should have a nozzle of proper size
and should provide sufficient pressure to completely fill
the joints. An extension for the nozzle of the caulking gun
and a longer tool for tooling the inner seal of a two-stage
joint are necessary. Joint filling should be done carefully
and completely, by thoroughly working the sealant into
the joint. Under-filling of joints normally leads to adhesion
loss. After joints have been completely filled, they should
be neatly tooled to eliminate air pockets or voids, and to
provide a smooth, neat-appearing finish. Tooling also provides a slightly concave joint surface that improves the sealant configuration and achieves a visually satisfactory finish.
Joint tooling should be performed within the allowable time
limit for the particular sealant. The surface of the sealant
should be a full, smooth bead, free of ridges, wrinkles, sags,
air pockets, and embedded impurities.
Large daily temperature swings during curing (warm days,
cold nights) may cause adhesive failure. A practical range of
installation temperatures, considering moisture condensation or frost formation on joint edges at low temperatures
and reduced working life at high temperatures, is from 40
to 80 °F (5 to 27 °C). This temperature range should be
assumed in determining the anticipated amount of joint
movement in the design of joints. A warning note should
be included on the plans that, if sealing must take place for
any reason at temperatures above or below the specified
range, a wider-than-specified joint may have to be formed.
Alternately, changes in the type of sealant to one of greater movement capability or modifications to the depth-towidth ratio may be required to secure greater extensibility.
The applicator should know the joint size limitation of the
sealant selected.
When it is necessary to apply sealant below 40 °F (5 °C),
steps must be taken to ensure clean, dry, frost-free surfaces.
The area to be sealed should be wiped with a quick-drying
solvent that is slightly water soluble, such as IPA, just before
sealing. The area may be heated, if possible, or at least the
sealant should be slightly warm (60 to 80 °F [15 to 27 °C])
when applied.
It is recommended that tools be used dry. Tooling solutions
such as water, soaps, oil, or alcohols should not be used
unless specifically approved by the sealant manufacturer as

374

|

ARCHITECTURAL PRECAST CONCRETE

they may interfere with sealant cure and adhesion and create aesthetic issues.
It is imperative that uncured silicone or polyurethane sealants are not allowed to contact non-abradable surfaces
such as polished stone, metal, or glass. These surfaces must
be masked or extreme care taken to prevent any contact
with the sealant during the application process. Excess sealant cannot be completely removed with organic or chlorinated solvents. Once an uncured sealant comes in contact
with an exposed surface it will leave a film that may change
the aesthetic or hydrophobic surface characteristics of the
substrate.
Surfaces soiled with sealant materials should be cleaned
as work progresses; removal is likely to be difficult after the
sealant has cured. A solvent or cleaning agent recommended by the sealant manufacturer should be used.
Sealant Backing. For sealants to perform to their optimum movement parameters, they must adhere only to the
joint sides and never to the base. Closed-cell expanded polyethylene, or non-gassing polyolefin sealant backing are the
recommended backing materials for horizontal and vertical
joints. For two-stage joints, open-cell polyurethane backing
should be used on the interior seal unless the interior seal is
allowed to cure for seven days before installing the exterior
seal. Proper selection and use of backing material is essential
for the satisfactory performance of watertight joints. When
selecting a backing material and/or bond breaker, the recommendations of the sealant manufacturer should be followed to ensure compatibility with the sealant.
The principal functions of sealant backing materials are:
1. Controlling the depth and shape of the sealant in the
joint (proper width to depth ratio). Also, profiles the rear
surface to an efficient cross-section for resisting tensile
forces.
2. Serving as a bondbreaker to prevent the sealant from
adhering to the back of the joint. The sealant must adhere only to the two surfaces to which it bridges. If it
also adheres to the back of the joint (three-sided adhesion), the stresses on the sealant bead are greatly increased and this increases the likelihood of premature
sealant failure.
3. Assisting in tooling of the joint by providing back pressure when tooling. The combination of tooling and
back pressure ensures full-sealant contact with the sides
of the joint, which is vital if proper adhesion is to take
place.
4. Protecting the back side of the sealant from attack by

DESIGN

4.7.7 Sealant Materials and Installation / 4.7.8 Architectural Treatment

moisture vapors trying to escape from the building. Use
of two-stage joints and backing is recommended where
high vapor pressure occurs at the immediate back surface of the sealant.
The backing should not stain the sealant, as this may
bleed through and cause discoloration of the joint. Sealant
backing materials should be of suitable size and shape so
that, after installation, they are compressed 25 to 50%.
Compression differs with open- and closed-cell rods; refer
to manufacturer’s recommendations. Adequate compression is necessary so that the shape will stay in the opening
and not be dislodged or moved by sealant installation.
Primers. Some sealants require primers on all substrates;
others require primer for specific substrates or none at all.
Absence of a required primer will cause premature sealant
adhesion failure. A primer often helps sealant adhesion in
cold weather. Primers are recommended by the sealant
manufacturer for the following reasons:
1. To enhance adhesion of sealants to porous surfaces,
such as concrete, or to reinforce the surface.
2. To promote adhesion of sealants to surfaces such as
porcelain enamel, unusual types of glass, certain metals
and finishes, and wood.
3. To promote adhesion of sealants to an existing surface
treatment which is difficult to remove.
Special care should be exercised to avoid staining the visible face of the precast concrete unit because some primers
leave an amber-colored stain if brushed along the surface.
This stain will have to be mechanically removed, which will
be expensive. The primer should be allowed to cure before
application of the sealant. Sealant must be applied the same
day the surfaces are primed. The sealant and primer should
always be supplied by the same manufacturer.

4.7.8 Architectural Treatment
Joints should be expressed as a strong visual feature of
architectural wall design. False joint lines can also add to
the visual effect. Recessing of joints and/or sealants will help
diminish the visual impact of possible variations between
adjacent surfaces sometimes inherent in large wall panels.
Setting the sealant back from the face of the panel also
gives some protection from UV light to minimize deterioration. By recessing the joints, the sideways flow of winddriven rain over the sealant is reduced. Complicated edge
and fenestration profiles should be avoided for economy
in manufacturing and erection. Complicated profiles are
more vulnerable to damage in handling and more difficult

4

to make watertight.
Joints are important features in creating weathering patterns. Vertical joints help in channeling water, provided the
joint is not pointed flush with a sealant or gasket. The concentration of water at such joints requires careful detailing
to prevent moisture penetration.
Listed are detailing suggestions for typical architectural
precast concrete panel joints (see Fig. 4.7.5, page 369).
1. Allow either a chamfered or reveal joint because these
types of joints can accommodate the tolerances required for panel thickness, and the shadows formed
within these joints will minimize any adverse effects on
the aesthetic appearance of the joint system. By making
the joints appear wider than they actually are, the visual
differences in their width are proportionately reduced.
This tends to make differences more difficult to detect
and masks slight misalignments of the joints that might
otherwise be especially noticeable at intersections.
Simplifying the profile of the joints by providing a reasonable radius (chamfering) the panel edges assists in
sealant installation and also has the obvious advantage
of making the edges less vulnerable to chipping. Chips
disrupt water flow and concentrate dirt.
2. Avoid the use of butt joints without a radiused or chamfered edge, as the tolerance variations in surface plane
may result in the formation of unwanted shadow lines
directly over the panels rather than within the joint area.
This may impair the aesthetic appearance of the panel
assembly.
Listed are detailing suggestions for staggering architectural
precast concrete wall panels (Fig. 4.7.7).
1. Check for excessive thermal bowing of panels and set
panel tolerances to avoid unwanted shadow lines at
certain times of the day.
2. Consider joint configuration and joint tolerances to
minimize unwanted shadow effects.
3. For loadbearing walls, there is a serious drawback to
using horizontally staggered panels. If staggered panels are used, the floor slab must bear on two different
panels on every other floor. The floor slab connection
problem created should be avoided, if at all possible.
Finish requirements may also influence joint details. The
sealant must be applied to a relatively smooth surface as
it is difficult to tool the sealant to achieve intimate contact
with an irregular surface. Thus, the sealant must be held

ARCHITECTURAL PRECAST CONCRETE

|

375

DESIGN

4.7.8 Architectural Treatment / 4.7.10 Joints in Special Locations

4.7.9 Fire-Protective Treatment
Joints between wall panels are similar to openings. Most
building codes do not require openings to be protected
against fire if the openings constitute only a small percentage of the wall area and if the spatial separation is greater
than some code minimum distance. In such cases, the joints
would not require protection. In other cases, openings, including joints, may have to be protected for fire resistance.
Where no openings are permitted, the fire resistance required for the wall should be provided at the joints.
Fire tests of wall panel joints have shown that the fire endurance, as determined by a temperature rise of 325 °F (163
°C) over the unexposed joint, is influenced by joint type, joint
treatment (materials), joint width, and panel thickness.
When required for fire rating, joints between wall panels
should be detailed to prevent the passage of flames and
hot gases. Details should ensure that the transmission of
heat through the joints does not exceed the limits specified
in ASTM E119 Standard Methods of Fire Tests of Building
Construction and Materials. Concrete wall panels expand
when heated, so the joints tend to close during fire exposure. By providing the proper thickness of insulating materials within the joint, it is possible to attain fire endurances
essentially equal to those of the panels. Flexible, noncombustible materials, such as ceramic fiber blankets, provide
thermal, flame, and smoke barriers. These fire resistive blankets and ropes must be installed with a minimum of 10 to
15% compression. When used in conjunction with caulking
materials, they can provide the necessary fire protection and
weathertightness while permitting normal volume change
movements. Joints that do not require movement can be
filled with mortar.
Figures 4.7.8 and 4.7.9 show the fire endurance of onestage joints in which the joint treatment consisted of sealants and polyethylene backer rods.
Table 4.7.2 is based on results of fire tests of panels with

376

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 4.7.8 Fire endurance of one-stage joints.
3
6 in. Panel
2
5 in. Panel
1
4 in. Panel
0

0

1/4

1/2
3/4
Joint Width, in.

1

Sealant

Backer Rod
Panel
Thickness

back 1/2 in. (13 mm) from the edge of exposed aggregate
and that portion of the matrix along the joint should present a smooth, clean surface for the application of the sealant (see Section 5.2.1). This requirement is simple to comply
with when the design includes recessed external joints (Fig.
4.7.5). When exposed aggregate surfaces come together
at an inside corner, the situation is more difficult. Special attention must be paid to surface finish and joint details. Also,
for maximum performance, sealants should not be applied
to beveled or chamfered surfaces, but should be applied
beyond the beveled area.

Fire Endurance, Hours

4

Joint Width

Fire Side
One-Stage Joint

one-stage joints and ceramic fiber felt in the joints. The tabulated values apply to one-stage joints and are conservative
for two-stage joints. Fire-resisting silicone sealants can provide fire ratings, if required. For high ratings, fire-retardant
joint filler materials may also be required.

4.7.10 J oints in Special Locations
Below-grade joints between panels and the foundation
require special attention. Good site drainage is essential for
long-term waterproofing. A perforated drain tile should be
placed below the top elevation of the floor slab. The top of
the drain should be covered or encased with a filter fabric.
The amount of coarse aggregate and its placement depend
on soil type, amount of groundwater expected, and depth
of the foundation. Where possible, slope the drain at least
1
/8 in./ ft (3 mm/300 mm), and close off the end with wire
mesh to keep rodents out. The discharge from drains should
be carried away from the foundation.
Specifying the proper backfill density for compacted soil
(between 85 and 88% on the Modified Proctor Density
scale) is extremely important. Density above 88% can in-

DESIGN

4.7.10 Joints in Special Locations

duce stress on the walls and impede drainage; density below 85% can result in some settlement.

The joint at the interface of the panel and foundation is
typically grouted and the grout is raked out on the earth
side and a backing material and sealant are installed. Due
to large variations in cast-in-place concrete foundations, a
minimum 1 in. (25 mm) joint is recommended at this interface. Damp-proofing materials may be used in the absence of hydrostatic pressure to resist the capillary action of
moisture. Damp-proofing should be stopped below grade
because, if exposed above a receding grade, it becomes a
visible black line.

Fig. 4.7.9 Exterior joint sealing configurations and fire
ratings (per UL 263).
11/2"
1"

Silicone
Building Sealant

Backer Rod

Wall

The amount of hydrostatic pressure expected in a building application also can be critical in material selection. For
most buildings, hydrostatic pressure around the foundation
is not a crucial factor, particularly if tile drains are installed
and working properly.

Wall

11/2-hour Fire Rating

However, a thorough analysis of groundwater levels and
soil percolation rates surrounding the site should be made
before deciding on the use of damp-proofing in lieu of
waterproofing.

1-3"
/ " Minimum

1 2

Silicone
Building Sealant

The joint between foundation panels should be caulked
and then covered with a sheet waterproofing membrane,
20 to 120 mils (0.50 to 3 mm) thick. The entire foundation
wall should then be covered with a sheet waterproofing
membrane and an asphaltic protection board or grooved
extruded polystyrene board with applied geotextile fabric.

5" Concrete Wall
(Minimum)

5" Concrete Wall
(Minimum)

4

3" Mineral Wood
/ " Minimum

1 2

Two-hour Fire Rating

Panel
Thickness*
(in.)
4

Thickness of ceramic fiber felt (in.) required for
fire resistance ratings and joint widths shown
Joint width = 3/8 in.

Joint width = 1 in.

1 hr 2 hr 3 hr 4 hr 1 hr 2 hr 3 hr 4 hr
/4

1

5

0

6
7

N.A. N.A. N.A.

/4

N.A. N.A. N.A.

/2

21/8

N.A. N.A.

/4

11/4

31 /2

N.A.

/8

2

33/4

3

/4

N.A. N.A.

0

0

11/8

N.A.

1

0

0

0

1

1

3

1

/4

7

Joint Width

Fig. 4.7.10
Sealant
Backer Rod
11/4 in. Ceramic
Fiber Blanket

Panel
Thickness

Table 4.7.2 Protection of joints between wall panels utilizing ceramic fiber felt.

Bond Breaker
Sealant
Fire Side
Vertical

Sealant

N.A. = Not applicable

The tabulated values apply to one-stage joints and are conservative for two stage
joints. Interpolation may be used for joint widths between 3/8 in. and 1 in.

Backer Rod
Ceramic
Fiber
Blanket

Note: 1 in. = 25.4 mm.

Panel
Thickness

* Panel equivalent thicknesses are for carbonate concrete. For siliceous aggregate concrete change “4, 5, 6, and 7” to “4.3, 5.3, 6.5, and 7.5”. For sandlightweight concrete change “4, 5, 6, and 7” to “3.3, 4.1, 4.9, and 5.7”

Joint Width
Fire Side
Vertical or Horizontal

ARCHITECTURAL PRECAST CONCRETE

|

377

5

The Paramount,
Architect: Elkus Manfredi Architects, Ltd.,
Design Architect; Kwan Hemmi Architects
and Planners, Architect of Record.

St. Regis Museum Tower,
Architect: Skidmore, Owings & Merrill, LLP.;
Photo: Mark Schwettmann.

San Francisco Museum of Modern Art,
San Francisco, California; Architect: Mario
Botta, Design Architect; Hellmuth, Obata &
Kassabaum, P.C. Architect of Record.

CHAPTER FI VE
OTHER ARCHITECTURAL
DESIGN CONSIDERATIONS
5.1 G
 ENERAL
Architectural precast concrete may be in contact with
many other materials under service conditions. Because
the application or interfacing of such materials can be as
important as the design of the individual components,
this chapter discusses various types and properties of
interfaces, including glazing, energy conservation and
condensation control, acoustical properties, fire and
blast resistance, and roofing. Joints between precast
concrete panels and other systems or materials must be
designed to maintain continuity of the thermal, air, and
moisture control functions in order to provide continuity for the wall. The combination of building materials
selected must provide an aesthetically pleasing facade
while effectively separating the external and internal
building environments. A building envelope is composed
of the architectural precast concrete panels, joints, and
other building materials discussed in this chapter.

5.2 W
 INDOWS AND GLAZING
5.2.1 D
 esign Considerations
For the architect, windows interrupt wall systems
and require special detailing at their interface with the
opaque cladding system. Windows are locations where
many dissimilar materials and different trades have to
interface.
Windows also affect the indoor environment because
their thermal and optical characteristics, size, orientation, ability to be opened, and treatment with shading
devices have an impact on the comfort of occupants
and the operation of mechanical systems. Glare, overheating and cooling effects, drafts, and outside noise
affect the physical comfort of occupants, while their
psychological comfort can also be affected by communication—or lack of it—with the outside world.
How much glazing, what type, and where to place it
to satisfy the occupants should be accounted for in the
evaluation of design options.
The fenestration industry has established guidelines for
bow, squareness, corner offset in window framing, and
variation of mullions from plumb or horizontals from
level:
1. B
 ow: 1/16 in. (1.6 mm) in any 4 ft (1.2 m) length of

framing.
2. Squareness: max. 1/8 in. (3.2 mm) difference in the
lengths of the diagonals of the frame.
3. Corner offset: 1/32 in. (0.8 mm) at each corner.
4. Plumb or level: 1/8 in. (3.2 mm) in 12 ft (3.7 m) or
1
/4 in. (6.3 mm) in any single run.
The fenestration industry has also established guidelines for the allowable deflection of glass framing members under design loads. The intention of the guidelines
is to minimize the potential for glass breakage. They
state that the glass surround should not impose any
bending or highly concentrated compressive loads on
the glass and that the framing should not deflect more
than 3/4 in. (19 mm) or 1/175 th of its span under loading, whichever is less. The guidelines also require that
the glazing system chosen should isolate the glass from
the other parts of the wall. The glazing option selected
should relate to realistic and attainable tolerances for
the precast concrete (see Section 4.6). For example,
erection of spandrel panels requires close attention to
tolerances because the window system must fit between the upper and lower concrete spandrel panels.
Connection of the window to the rough opening in a
precast concrete wall must be designed to resist wind,
seismic, blast, vertical live load, and thermal loads and
to transfer loads into the surrounding framing.
The interfacing between windows and precast concrete
panels is fairly simple in the case of closed shapes, where
the window is framed within a single panel. In this case,
connections and joint details are independent of site conditions and tolerances are governed only by tolerances
related to the manufacturing for the two products (Fig.
5.2.1). In the case of open units and spandrel panels,
the interfacing of windows in the façade must allow for
slightly larger and more uncertain site construction tolerances. Where window openings occur between such
units, glazing can best be accommodated by a window
frame, that considers the appropriate size and squareness
tolerances of the opening. Preferably, windows should be
located entirely within a single panel. Failure of the joint
where the window frame crosses the joint will result in
water being directed into the window head. A precast
concrete piece that fully contains the window opening
makes the most economical wall unit.
ARCHITECTURAL PRECAST CONCRETE

|

379

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.1 Windows and Glazing Design Considerations

exposure of rain water running down the face of
the building.

Fig. 5.2.1 Open and closed units.
Joint

Open

Joint

Open

Joint

Closed

It is important that panel-to-window connections for
open units and spandrels allow for minor movements
of the panels in relation to the supporting structure
and in the joints between the concrete and window
frame. Consideration should be given to the effect of
gravity and seismic deflections and rotations of the
spandrel on the windows. In the case of closed panels,
the movements are accommodated in the joints between the concrete units.
One or two lines of sealant can be used as the means
of providing a watertight joint between the window
and the precast concrete wall. The window joints at the
precast concrete panel interface have leaked due to unsatisfactory surface preparation or improper sealant selection leading to cohesion or adhesion failure. Leakage
has led to moisture damage to ceilings, floor and wall
finishes and structural elements and reduced the insulating capacity of wall insulation attached to the back of
the precast concrete panel. Using an appropriate water
penetration test as part of the testing program for the
mockups and installed window units, along with proper detailing and installation procedures, helps protect
against water leakage problems in service.
In the design of the building façade, there are practices that can be considered that have a proven history
of deterring leakage at windows. These basic design
practices are often overlooked, or their effectiveness
underestimated. Some of these practices include:
• Roof overhangs or cornices – These features can reduce direct exposure to rain. Effectiveness relates to
the degree of overhang protection, site exposure,
and other factors.
• Drips – Incorporated into panels they can shed water away from window joints, glazing, and seals.
• Recess the window – Even a small degree of recess
into the wall plane affords some protection from

380

|

ARCHITECTURAL PRECAST CONCRETE

• Slope sills – Slope sills away from windows. Avoid
creating horizontal or back-sloping surfaces that
will allow water and snow to collect at sills.
It is important to detail the extent and location of
precast concrete finishes in the contract documents,
particularly when an exposed aggregate or washed
texture is indicated. The reason being the difficulty in
achieving a satisfactory face seal of the window frame
to exposed aggregate. Pin holes at the sealant bond
line will be created by the irregular exposed aggregate
surfaces since the sealant cannot be effectively tooled
and consolidated. The most effective means to avoid
pinholing is to hold the texturing back a minimum of
1
/2 in. (13 mm) from the precast concrete and window
frame interface. Light sandblast, acid-etch, and other
smooth textures are less problematic because a suitable and continuous bond to the precast concrete can
be achieved by the sealant.
The specifier should also ensure compatible sealants
are used between the precast concrete and window
systems. One means is to specify installation of all exterior wall sealants by a single sealant applicator. Often,
a joint sealing system shows a lack of continuity by failing to include the window perimeter. The interface of
the two systems requires a “marriage” bead of sealant
where the vertical precast concrete joint sealant meets
the horizontal window sealant.
Water drips. Surface tension allows water to flow
along the underside of horizontal surfaces. If no drip
devices or projections are provided on a building face,
run-off water can flow over the wall materials and windows for the total height of the building. Dirt may be
deposited in sufficient quantities to cause stains on wall
units and, in a short period, may streak and stain glass
(Fig. 5.2.2) (see Section 5.2.4). Water always runs down
a wall or window over the same preferential paths.
Window heads should be designed so they don’t splay
down and back toward the glass, unless drip details are
incorporated into the frame. Water drips will minimize
streaking due to uneven washing of a backward-sloping surface when the drips are correctly dimensioned
and placed close to the forward edge. When placed
under outward-sloping heads or sills, drips will reduce
streaking on the vertical surfaces below. These drips
also prevent water (after a storm) from slowly running
over the window glass, a primary cause of glass streak-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.2.1 Windows and Glazing Design Considerations

5

drip should not be located closer than 11/2 in. (38 mm)
to the edge of the precast concrete unit. Where the
window is not 2 in. (50 mm) or more back from the
face of the panel, it is difficult to get a drip groove in
the panel.
A clear sealant bead applied to precast concrete units
after erection, or plastic drips glued to the concrete,
are remedial drip solutions used with varying success
depending on their care in application. A drip incorporated initially into the precast concrete or window
frame is the least-costly and best solution. Figures
5.2.5(a) and (b) show the use of an extrusion [either
aluminum or neoprene across the head of the window
which have either an integral gutter or extended drip
lip of at least 1 in. (25 mm)].
Fig. 5.2.4 Design of water drip in relation to slope.
Fig. 5.2.2

ing. Water will leave a drip at its lowest point and it is
important to follow its course thereafter. Small chips
and cracks in the building surface may concentrate the
flow, so that water will bridge drip details and allow
wetting of the surface below. If particulate laden water
falls onto other surfaces, the problem may be merely
relocated. However, if the wind tends to spread the
water out on the surface below, uniformity of weathering may be obtained. To avoid streaks on the sides of
window panels, the drip may be stopped about 2 in.
(50 mm) short of the window sides (Fig. 5.2.3). Often
recesses or grooves may be incorporated in the side
walls to further direct the water. Water flow should be
evenly distributed on wall surfaces.
The drip section should be designed in relation to the
slope of the concrete surface (Fig. 5.2.4) to prevent
water from bridging the drip. To avoid chipping, the
Fig. 5.2.3 Straight water drip.

11/2 in. min.
(38 mm)

2 in. min.
50 mm

Sealant or plastic drip

Flashing. One-stage (face-sealed) sealant joints,
which are most common, two-stage joints, or flashing
may be used to prevent moisture intrusion. Flashing is
not used in most regions of the United States. Its use
is dependent on the design of the window system, design wind pressure, and expected annual precipitation.
Improperly detailed and installed flashing may cause
water leakage problems.
Weep Holes. The glazing guidelines published by
most major window associations recommend incorporating weep holes at the sill of the glazing pocket for
the control and management of infiltrating water and

ARCHITECTURAL PRECAST CONCRETE

|

381

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.1 Windows and Glazing Design Considerations

Panels incorporating punched openings produce more
accurate opening sizes than ribbon windows created
by a column and spandrel system. This can be an important factor if the glass units are preordered.

Fig. 5.2.5 Gutter or drip incorporated in drip section.

(a)

Window openings can be provided in architectural
precast concrete panels with ease, in any shape or size
desired, offering the architect total flexibility in design.
Panels may contain a single opening or a series of windows. They can be one-story high and made as wide
as possible or cast narrower to span vertically for two
floors. However, achieving design and cost efficiency
requires thoughtful panel system configuration. In all
cases, a reasonable slope must be maintained on the
return edges of openings to ensure sufficient draft,
usually 1:6, to strip the unit out of the mold. Mold
costs are directly related to the complexity of the window wall panel. Punched flat panels add moderate
costs, whereas the molds for heavily sculptured panels can be expensive. As always with precast concrete,
repetition of components reduces unit costs.
The designer should consider window wall panels that:
• Promote the use of a master mold.
• Offer flat or heavily sculptured profiles.
• Provide curved surfaces.
• Work as corner units.
• Incorporate a bullnose, cornice, or reveals.

(b)

1 in.

condensation. The need for weep holes is a function
of how the system is designed. Modifying pretested
window systems solely to add weep holes is not recommended. Unless the window system is designed
properly, weep holes can be a source of water intrusion
and increased air infiltration.
Window Frame Location and Detailing. Special
consideration should be given to the relationship between window frames and architectural features, such
as reveals and projecting elements on a precast concrete panel. For example, window installers discourage
aligning the frame’s exterior with a series of reveals.
A design that ensures installation success features
window returns that create a smooth surface against
which the installer can set and plumb the frames.

382

|

ARCHITECTURAL PRECAST CONCRETE

Openings in walls that form the windows of a building help to break up large flat surfaces, but may contribute weathering problems. Individual windows in a
wall made of architectural precast concrete must be
designed with two principles in mind: (1) to contain
the water flow within the window area, and (2) to disperse it at the bottom of the window in such a manner
that it is spread not concentrated.
Glass areas cause build-up of water flow. Typically
on nonabsorbent materials, water flows in discrete
streams rather than as a continuous film. This is due
to surface tension that causes the droplets of water to
converge.
Sills should have a minimum slope of 2% to ensure
water flow away from the interior of the structure
and minimize dirt accumulation. A drip groove should
be provided under any outward-sloping sills to prevent particulate laden water stains on the wall panels. Because window surfaces are impervious, a large
amount of the water and pollutants that collect on

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.2.1 Windows and Glazing Design Considerations

Fig. 5.2.7 Window treatments.

Exterior

windows are flushed over the sills, the very thing the
designer must guard against if differential patterning is
to be avoided. In most cases, sills should be designed
to allow rainwater to pass over windows, sills, and
spandrels as evenly and freely as possible. By contrast,
rainwater flowing down an adjacent precast concrete
wall surface will be slower (depending on the surface
texture and absorption of the concrete) and its throwoff will be less complete. As a result, a concentration
of water forms at the bottom of each window, which

5

(a)
Textured or
profiled concrete

(b)
Coffered

(c)
Deep recess

may cause differential patterning. This flow must be
dissipated, breaking up its concentration. Furthermore,
there is always a tendency for more water flow at the
edges of the glass or at the mullions than in its center.
This effect is due to small amounts of negative wind
forces which tend to drive rain toward the edges of
the glass. Figures 5.2.6(a) and (b) shows the dirt pattern caused by water run-off carrying the dirt down
the mullion. Grooves were cut under the mullions (Fig.
5.2.6[a]) in an attempt to minimize staining after the
panels were in place.

(a)
Fig. 5.2.6(a) & (b)Water flow over glass depositing dirt.

(b)

There are many different ways of detailing windows,
depending to some extent on their shape and the degree to which they interrupt water flow (Fig. 5.2.7).
Many recent buildings have used precast concrete
spandrel panels and windows in approximately the
same plane, without sill ledges or projections, and
the panels have weathered well without staining. This
method seems to perform best when the concrete below the window is textured or profiled (for example,
with vertical ribs) to break up the water flow and avoid
streaking. Darker color finishes are also desirable in urban areas. The path of water flow must be anticipated
and provision made to collect and drain away the water in due course. See Section 3.6 for a discussion on
weathering problems.
Important design considerations are the details for
handling of water directed to the base of a window or
glass curtain wall. If the water volume is considerable
(tall sections of glass) and the glass is not kept clean,

ARCHITECTURAL PRECAST CONCRETE

|

383

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.1 Windows and Glazing Design Considerations / 5.2.2 Window Installation

a large volume of particulate water will be directed to
the sill and then the precast concrete panel surface.
The dirt is more noticeable if the under-window panels
are made with light-colored concrete. Ideally detailing
should be such that water flow from the base of a window is thrown clear of the building, by setting the precast concrete wall behind the line of the mullions and
curtain wall sill. As alternatives, water may be made
to flow in rustications in line with the mullions for
the glass. Shadows on rustications usually help mask
streaks, particularly when the recessed depth is equal
to or greater than the recess width.

5.2.2 W
 indow Installation
Precast concrete panels occasionally are pre-glazed in
the precaster’s plant (Fig. 5.2.8). In the project shown,
the precaster and glass subcontractor coordinated operations to pre-glaze the punched window panels. On a
just-in-time basis, panel loads were shuttled to the glass
subcontractor where the glass was installed on the
panels–right on the delivery trailer. When glazing was
completed, the loads were re-delivered to the jobsite
for erection on a 42-story tower. During the whole operation, including erection, only one or two panes were
lost. Pre-glazing the precast concrete panels offered
significant cost and schedule savings on this project.
However, the typical method of window frame attachment is field installment. This is accomplished with
expansion anchors in the edge or back vertical face of
the panel. Location for window attachments should be
coordinated with the precaster to avoid interference
with reinforcing steel or prestressing tendons. Proper
edge distance for the anchors must be used. Impact
type anchors should never be used.
Hardware can also be installed in the precast concrete
units to provide fastening for the windows and may
consist of ferrule loop inserts, tubes or slotted inserts.
Several methods are commonly used to attach windows to precast concrete panels. In precast concrete
sandwich wall panels, the window frame should be
attached to the inside wythe, since movement of the
inside wythe is less than the exterior wythe. Window
frames should have thermal barriers between the exterior frame and the interior frame. A substantial part of
the total heat loss through a window can occur through
its frame. Windows require careful detailing at their interface with the wall. Consideration needs to be given
to: (1) mechanical connection of the window frame to

384

|

ARCHITECTURAL PRECAST CONCRETE

concrete panel; (2) estimation of joint movement for
proper selection of sealing materials; (3) watertightness; (4) airtightness; (5) flashings, rain deflectors, and
drainage of infiltrated water; and (6) condensation.
Figures 5.2.9(a) through (f) show several examples of
aluminum frames attached to precast concrete panels.
Figure 5.2.10 shows 7 ft 3 in.-high (2.2 m) loadbearing
spandrel panels with butt-glazed insulated glass and
spandrel glass partially covering the spandrel panels.
The GC/CM must coordinate the supply of window
hardware that is typically embedded in the precast
concrete. This hardware is frequently furnished directly
to the precast concrete plant by the window supplier.
If ferrous metal inserts are used for aluminum windows, they should be galvanized, coated with two coats
of bituminous paint or zinc-rich primer. Depending on
the finish of the aluminum, plastic washers may have
to be used to prevent contact between the steel and
the aluminum to prevent possible corrosion.
Where windows are installed using a two-stage joint,
it may be necessary to vent the air space between the
rain-barrier and the airseal (refer to Section 4.7.2).
The amount of water to be drained should normally
be minimal so that ice lensing or damming problems
will not occur, provided the outlet is protected with a
baffled vent or weep tube.
Untested window frame to precast concrete assemblies should be mocked up and water tested in accordance with ASTM E 331 during the initial precast
concrete production stages. This allows any necessary
corrective action to be taken at the earliest possible
time. It is important to coordinate the window sealant
system with the adjacent precast concrete panel sealant system.
Fig. 5.2.8 Pre-glazing of panels.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.2.2 Window Installation

5

Fig. 5.2.9 Window attachment details.

Note: Window system not
attached at jamb

Head

Head

(b)

(a)

(c)
Jamb

Sill

Sill

(f)
(d)

Head

Sill

(e)

Head

Head

Sill

Sill

ARCHITECTURAL PRECAST CONCRETE

|

385

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.2 Window Installation / 5.2.3 Other Attached or Incorporated Materials

2"

11/2"

Fig. 5.2.10 Loadbearing spandrels partially covered with glass.

Sealant with
Backup Rod
/"
2/"
1 2

1 4

Cont. Horizontal
Alum. Mullion
8"

Precast
Concrete Panel
Cont. Shim
as Required

/ " Spandrel Glass

1 4

Alum. Clip Anchor

1"

21/4"
Extruded Alum.
Head Trim

Alum. Frame with
1" Insul. Glass

Mini Blind
Window Cover
63/4"

Window Frame Location. When possible, it is preferable to locate the window frame 2 in. (50 mm) or
more from the exterior face of the panel. This helps
reduce the chance of glass staining by providing space
for a drip and allows lifting devices to be concealed
beneath the window frame. As window frames are
punched back in the opening return consideration
must be given to achieving the return finishes.
It is recommended that the architect specify, as part
of the glazing contract, that the window manufacturer
install window(s) for approval in the initial production
panel or previously approved mockup units. This practice often proves advantageous because it affords the
architect and the owner an opportunity to assess the
appearance of the finished wall assembly and to approve all details relating to the assembly such as color
and type of sealant and interior trim. The mockup sample may also allow subcontractors to pinpoint problems before large numbers of elements are affected.
This provision is a worthwhile safeguard against later
delays in the enclosure of the building.

386

|

ARCHITECTURAL PRECAST CONCRETE

5"

5.2.3 Other Attached or
Incorporated Materials
Other materials incorporated into the precast concrete
units might include inserts or hardware used to attach
other materials, reglets used to accommodate flashings, nailing strips, or similar continuous fastening strips.
Materials which react with concrete should not be used.
For example, metals such as copper, zinc, aluminum and
lead, and alloys containing these metals may corrode
when in concrete. If their use is unavoidable, they must
be suitably protected with dielectric separation materials.
Galvanized and stainless steel and plastics are acceptable
without dielectric separation but, as indicated in Section
3.6.3, the weathering effects of these and other metals
should be considered. Also, flashings must be galvanically
compatible with the reglets or counterflashing receivers.
There has been great progress in window-washing
systems. More sophisticated safety equipment has
been designed into the systems and more flexibility has
been given to the designer of the wall.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.2.3 Other Attached or Incorporated Materials / 5.2.4 Glass Staining or Etching

Window-washing tiebacks prevent the windowwashing scaffold from swaying or separating from the
wall (Fig. 5.2.11) and lend themselves to designs that
are flush, with no projections in the façade.
It is extremely important for the architect to locate
attached or incorporated materials on the Contract
Drawings and clearly indicate the supplier and installer
of these materials in the specifications.

5

Fig. 5.2.11 Tieback.

Window-washing tieback

5.2.4 G
 lass Staining or Etching
Glass corrosion is a gradual wearing away of the glass
by chemical action. Corrosion may occur when any of a
variety of compounds, such as cleaning solutions containing ammonium bifluoride, contaminate the surface and form highly acidic or alkaline solutions when
mixed with water. These compounds may come from
the atmosphere in the form of hydrocarbons or pollutants, or they may come from within the glass itself. If
the material in contact with glass is inert and moistureproof, the material will protect the glass surface from
changes caused by exposure to moisture. Later, if the
contacting material is removed, a differential surface
change may become quite visible and unattractive
under some lighting and viewing conditions, even if
the change is slight. In addition, some silicone sealants have ingredients, usually plasticizing oils, that may
leach out and combine with other airborne particles
to stain the glass. Also, primer or silicone sealant that
overlaps the joint onto glass will leave a permanent
mark because the silicone will chemically bond with
the glass. Careful masking of the sealant joint during
installation can prevent this type of discoloration.
When a small volume of water maintains contact
with a relatively large area of soda-lime-silica (sodium
calcium silicate) glass, conditions may be conducive to
leaching of the alkaline materials in the glass. The interaction of glass and water results in the replacement
of the sodium ions in the glass with hydrogen ions
from the water. As sodium ions accumulate and hydrogen ions decrease in a thin film of water, the liquid
will increase in alkalinity at a much greater rate than
if it were absorbed into a large volume of water. Also,
this reaction and, the solution pH increases much more
rapidly at elevated temperature (140°F vs. 73°F [60°C
vs. 23°C]). As long as the alkaline concentrations of
the resulting solution remain below pH level of 8.5 (the
threshold of permanent surface damage), glass etching does not occur. However, if the evaporation rate is

very slow, and the pH level increases to 9.0 or above,
glass network dissolution (glass etching) displaces ion
exchange as the predominant reaction mechanism.
This introduces glass ions such as calcium, magnesium,
sodium, and silicon into solution.
Glass-surface corrosion requires a moist, stagnant
environment in which water remains in prolonged
and undisturbed contact with glass. Any factor that
increases the length of time moisture is in contact with
glass during wet-dry cycles is likely to speed tenacious
staining and possibly contribute to glass etching. One
of the most common contributors of differential wetting is dirt or dust. Dirt accumulation on glass holds the
water on the glass longer causing moisture to attack
the surface. The finely divided damp materials in contact with glass cause the glass constituents to dissolve
slightly and be redeposited at the evaporating edge of
the material, resulting in tenacious deposits. (Glass will
normally lose some of its sodium ions by dissolution in
water, but the calcium ion then stops most of the dissolution. In polluted atmospheres, the acids—SOx or
NOx —will attack the calcium ion and permit further
dissolution.) If the dissolved material is washed away,
little change can be seen by the human eye. But when
the solution remains on the glass, atmospheric carbonation of the alkali metal and alkaline earth silicates
causes a subsequent tenacious deposit of silica gel.
Upon evaporation at a temperature of 73°F (23°C),
dilute solutions of silica and silicates, with a concentration as low as 4 to 8 ppm (this includes hard water),
can reach a point at which super-saturation occurs.
This is accompanied by the rapid formation of polymerised silica and polysilicate species that generate
amorphous, water-insoluble precipitates of silica gel
on the glass. The silicates form chemical bonds to the

ARCHITECTURAL PRECAST CONCRETE

|

387

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.4 Glass Staining or Etching

glass surface producing a “glass-on-glass” deposit.
This deposit, after aging and exposure to atmospheric
acids resists conventional cleaning agents and usually
requires removal using polishing compounds such as
optical grade cerium oxide or 4F pumice.
The hardened gel retains moisture and guides run-off
water along the same paths. The process of surface
corrosion then becomes self-perpetuating. When this
happens uniformly, the eye does not detect the differences. However, the silica gel deposit or the glass etch
depth need not be thicker than a wavelength of light
for the eye to detect it. Frequent washing of the windows tends to remove the gel before it becomes hard,
minimizing staining and etching of the glass.
Directed slow water run-off and the resultant dirt
accumulation, cause the glass to be attacked nonuniformly. Eventually, the cycle of water drying, gel
forming, acid atmosphere attack, and alkali-washing
compounds causes in-depth glass dissolution. Then,
the cost of cleaning and buffing can approach the cost
of window replacement. Thus, architectural design
and detailing including the use of drips can prevent
this condition from occurring.
Staining will be more noticeable on tinted heat-absorbing glass because of the greater contrast between
the lighter color of the stain or etch and the darker
color of the glass. In addition, heat absorption will increase the rate of etch. There is no known difference
in the composition of tinted glasses, which contributes
to this staining, as compared to clear glass. Staining is
also more noticeable on reflective glass.
The usual explanation for the etching of glass in concrete structures is that concrete contributes alkaline materials to the run-off water. Hydration of cement results
in the formation of hydrated calcium silicates (CSH),
Ca(OH)2, and aluminates, and the remaining water in
the concrete becoming highly alkaline. It is well known
that alkali (OH-), meaning high pH material, will attack
glass. What is not well known is that atmospheric acids
(NOx, SOx, and CO2) can quickly neutralize low concentrations of these alkalies from concrete to produce
less alkaline salts of calcium, sodium, or potassium. Of
these reaction products, only the carbonates of sodium
and potassium provide the most soluble alkaline salts.
However, even these salts are quickly converted to the
bicarbonates that are only very weakly alkaline.
Because the atmosphere is usually very acid in the

388

|

ARCHITECTURAL PRECAST CONCRETE

larger cities (refer to Fig. 3.6.1 in Chapter 3), low concentrations of leached alkali (high pH) will be neutralized. However, in the case of concrete (that is less than
28 days old), higher concentrations of alkaline leachate may not be completely neutralized before it has
reacted with the glass.
Although laboratory studies have demonstrated
that glass can be susceptible to alkaline-induced surface damage, it does so under conditions that do not
prevail in the environments typically encountered by
glazing systems. In these studies, a solution of calcium
hydroxide (pH 11.5) placed on glass at 140°F (60°C) in
a controlled environment to retard evaporation does
not cause chemical erosion or etching after 20 hours.
In the field, it is only the last water droplets after a
rain, adhering to glass, that could present a threat to
surface quality via alkaline etching, if indeed the solution pH is 9.0 or greater and the residence time is
well beyond 24 hours. It is doubtful that these droplets
could exist intact for the periods required for severe
alkaline etching to develop. Repeated deposition and
evaporation can eventually lead to tenacious deposits
and subsequent chemical etching.
Not all rainfall that contacts concrete has time to permeate the surface micro-pores to extract alkaline materials; in fact, much of the water flows away. Rainfall
can permeate concrete having high absorption and
cause efflorescence (see Section 3.6.8). The efflorescencing salts are usually neutralized by the carbon dioxide in the air within a relatively short period of time.
The leaching of alkali from most concretes becomes
dramatically reduced in one to two years because the
surface lime is mostly carbonated in place, and the interior alkalies within the concrete matrix cannot usually
be reached.
In addition, chemical reaction of the cement compounds with sulfur and nitrogen oxides in the air occurs with the subsequent precipitation by evaporation
of solutions containing the reaction products, such as
gypsum (CaSO4•2H2O). The transference to and deposition of these materials on the window glass by rainwater can result in surface staining and etching, if they
are allowed to remain on the glass for a period of time.
(The gypsum acts in the same way as dirt in causing a
stain.) The time period for a stain to result depends to
some degree on the ambient temperatures with warmer temperatures causing the stain to occur sooner.
The plasticity of fresh concrete lends itself to use in

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.2.4 Glass Staining or Etching

many shapes that may not incorporate proper water
run-off in the final, hardened design. In addition, the
rough surface textures of exposed aggregate concrete
increase water retention, which results in slow water
run-off. When uniform wetting of windows occurs,
staining of glass generally does not occur. However,
when differential wetting of the windows occur from
a slow run-off of rainwater, such as by dripping, stains
can occur regardless of the construction material used
above the windows.
Weathering steel, bronze, limestone, aluminum, or
precast concrete façades all may experience staining of
window glass (Fig. 5.2.12). Analysis of powder scraped
from glass stains on both a metal-clad and concreteclad building show that a good portion of the stain
is composed of gypsum indicating that SOx from the
atmosphere also plays a role in staining. The calcium

5

on the metal building had to come from airborne substances or from the glass.
Considerations for Prevention of Staining. Newly
developed hydrophobic coatings can be applied to
the glass surface to protect it from chemical attack.
However, there may be a limitation to the size of the
glass units that can receive the coating. Hydrophobic
coatings create a barrier between the glass and water and other pollutants. They help prevent the initial stages of corrosion and reduce the occurrence of
other materials like mineral deposits from sticking to
the glass. Hydrophobic substances repel water and
harmful chemicals and keep glass free from staining.
Microscopic-level glass surfaces are jagged with hills
and valleys and hydrophobic coatings fill the voids,
creating a smooth seal over the glass surface. This
smoother surface is more scratch-resistant, repels wa-

Fig. 5.2.12

ARCHITECTURAL PRECAST CONCRETE

|

389

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.2.4 Glass Staining or Etching / 5.3 Energy Conservation and Condensation Control Glossary

ter, and provides less adhesion for unwanted materials.
The performance life of hydrophobic coating products
can vary from weeks to over five years; however, they
can be reapplied.
Building details can reduce the amount of water discharged to the glass. Concrete frames at window heads
should, wherever possible, be designed so that they do
not splay down and back toward the glass unless drip
details are incorporated into the frames. Without drip
details, a direct slow wash-down of the glass should
be anticipated. A drip directs run-off water away from
the glass. This will prevent many staining and etching
problems.
An important line of defense against slow run-off is
the introduction of edge drips and a second drip or gutter. This can be accomplished by having a cast-in drip in
the concrete (see Section 3.6.2) or by using extrusions
(either aluminum or neoprene) across the head of the
window that have either an integral gutter or extended
drip lip of at least 1 in. (25 mm) (Fig. 5.2.5).
Periodic window washing (every 90 days during construction and every 180 days thereafter depending on
dirt accumulation) is important in minimizing stains
and corrosion from occurring on glass. By doing this
cleaning, deposits will not have time to accumulate.
It is important that glass be cleaned, rinsed, and dried
with a clean squeegee following rain or other washoff conditions, particularly during building construction. Since it is costly to ask for more than one washing
during the construction phase, it might be advisable to
include a provision for at least monthly examinations
of the glass surfaces. Then, if dirt, dust, plaster, drywall spackle, grout, paint splatter, or other construction
refuse is found, it can be removed before permanent
damage occurs. (These materials can combine with
dew or condensation to form mild chemicals that may
etch or stain the glass.) Washing of the windows when
deposits are first noticed minimizes staining and etching of the glass. Care should always be taken to prevent
window cleaning compounds from being washed over
the precast concrete panels or the panel surfaces will
need to be cleaned. Acceptable cleaning procedures
are available from glass manufacturers and fabricators.
A mild soap, detergent, or window-cleaning solution should be used to clean glass surfaces. The glass
surface should then be rinsed immediately with clean
water and any excess water removed from the glass

390

|

ARCHITECTURAL PRECAST CONCRETE

surface with a squeegee. Detergents or other cleaners
should be tested to make sure that they rinse away
completely without leaving a film, which could hold
dirt. Also, maintenance workers should clean the sealant joint as well as other building surfaces; otherwise,
the dirt left in the joint may wash out over the clean
surface. Harsh cleaners and abrasives, particularly
those of an alkaline character, are not recommended,
particularly on reflective glass. If a light stain or etch
remains, it may be removed with a slurry of cerium
oxide and water or 4F pumice and water mixed to a
paste consistency and applied with a blocking pad. As
an alternative, 4F pumice plus Windex may be rubbed
on the glass, applying light pressure with a clean damp
cloth, using 10 to 20 strokes. However, if the etch is
already deep, the cleaning procedures will not produce
a uniform surface by removing the streak because it is
already in a different plane and reflects light at a different angle. Great skill is required in use of buffing
heads to avoid creation of “bulls-eyes” and other nonuniform surface effects.

5.3 ENERGY CONSERVATION AND
CONDENSATION CONTROL
5.3.1 Glossary
Albedo – Solar reflectance; see reflectance.
Dew-point temperature – The temperature to which
air must be cooled so that it cannot hold any more
water. This depends on the ambient temperature, relative humidity, and pressure. Below this temperature,
condensation will occur. The temperature corresponding to saturation (100% relative humidity) for a given
absolute humidity at constant pressure.
Building envelope – The components of building that
perform as a system to separate conditioned space
from the exterior.
Daylighting – Illuminating the inside of a building
using natural light from the sun, rather than electrical fixtures. Daylighting controls can be used to dim
or turn off lights along the building perimeter when
daylighting is prevalent.
Film or surface resistance (Rf) – The thermal resistance of a thin layer of air adjacent to the indoor or
outdoor side of a wall or other building component.
Subscripts “i” and “o” are usually used to denote indoor and outdoor surface resistances, respectively.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3 Energy Conservation and Condensation Control Glossary / 5.3.2 Energy Conservation

Heat capacity – A measure of the ability of a material
to store heat, defined as the amount of heat necessary to raise the temperature of a given mass by 1°F.
Numerically, the sum of the products of the mass per
unit area (unit weight) of each individual material in
the roof, wall, or floor surface multiplied by its indi2
vidual specific heat. Units are Btu/(ft ·°F).
Perm – A measure of passage of water vapor or moisture flow through a material; specifically, the mass
rate of water vapor flow through one square foot of
a material or construction of one grain per hour induced by a pressure gradient between two surfaces
of one inch of mercury. A perm is 1 grain per ([sq ft of
area][hr][inch of mercury vapor pressure difference]). A
grain is equivalent to 1/7000 lb.
Permeability, water vapor (µ) – The property of
a material that permits the passage of water vapor,
defined as the time rate of water vapor transmission
through unit area of flat material of unit thickness induced by unit vapor-pressure difference between two
specific surfaces, under specified temperature and humidity conditions. It is equal to the permeance of a unit
thickness, generally 1 in. of a material, or the product
or permeance and thickness. Permeability is measured
in perm inches. The permeability of a material varies
with barometric pressure, temperature, and relative
humidity conditions.
Permeance, water vapor (M) – The time rate of water vapor transmission through unit area of flat material or construction induced by unit vapor pressure
difference between two specific surfaces, under specified temperature and humidity conditions. Permeance
is measured in perms.
Reflectance – The ratio of the amount of light or solar
energy reflected from a material surface to the amount
that shines on the surface. Solar reflectance includes
light in the visible and ultraviolet range. For artificial
lighting, the reflectance refers to the particular type of
lighting used in the visible spectrum.
Relative humidity (RH) – The ratio of water vapor
present in air to the water vapor present in saturated
air at the same temperature and pressure.
Specific heat – The quantity of heat energy in Btu’s
required to raise the temperature of one pound of a
material by 1°F.
Sustainability – Development that meets the needs

5

of the present without compromising the ability of future generations to meet their own needs.
Thermal conductivity (k) – The time rate of heat flow
through a unit area induced by a unit temperature difference between two defined surfaces of a unit thickness of a homogenous material under steady-state
conditions. Units are Btu·in./hr·ft2·ºF.
Thermal mass – A characteristic of concrete materials
with mass heat capacity and surface areas capable of
affecting building heating and cooling loads by storing and releasing heat as the interior and/or exterior
temperature and radiant conditions fluctuate. Use of
mass walls as part of the building envelope and in passive solar design reduces peak and total energy loads
for many buildings and climates. Steady-state methods
of measuring or predicting heat flow or energy use do
not take into account the dynamic effects of thermal
mass.
Thermal resistance (R) – A measure of resistance to
heat flow of a material, defined as the reciprocal of the
time rate of heat flow through a unit area induced by a
unit temperature difference between two defined surfaces of a material or construction under steady-state
conditions. Units are hr·ft2·ºF/Btu.
Thermal transmittance (U) – A measure of heat flow
through a material. The inverse of total or overall thermal resistance (RT). Units are Btu/hr·ft2·ºF.

5.3.2 Energy Conservation
Americans spend almost 90% of their time inside
buildings. More than 2/3rds of the electricity generated and 1/3rd of the total energy (including fossil fuels and electricity) in the U.S. are used to heat, cool,
and operate buildings. Significant energy could be
saved if buildings were built to meet or exceed minimum national energy code standards. Saving energy
will result in fewer power plants and natural resources
being used to provide electricity and natural gas. It
also means fewer emissions to the atmosphere. These
emissions have been associated with the formation of
smog, acid rain, and global climate change.
Energy codes provide minimum building requirements
that are cost effective in saving energy. The U.S. Energy
Conservation and Production Act (ECPA) requires
that each state certify that it has a commercial building code that meets or exceeds ANSI/ASHRAE/IESNA

ARCHITECTURAL PRECAST CONCRETE

|

391

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation

Standard 90.11. In this sense, “commercial” means all
buildings that are not low-rise residential (three stories
or less above grade). This includes office, industrial,
warehouse, school, religious, dormitories, and highrise residential buildings. Some states implement codes
similar to ASHRAE Standard 90.1 and some have other
codes or no codes. The status of energy codes by states
is available from the Building Codes Assistance Project
(BCAP) (www.bcap-energy.org/backissues.html). The
designer is not constrained in aesthetic expression
in applying the range of available high performance
building systems to meet the performance criteria of
ASHRAE 90.1.
Sustainability or green-building programs such as
LEED™2or EnergyStar3 encourage energy savings beyond minimum code requirements. The energy saved
is a cost savings to the building owner through lower monthly utility bills, and smaller, thus less expensive heating, ventilating and, air-conditioning (HVAC)
equipment. Less energy use also means fewer emissions to the atmosphere from fossil fuel power plants.
Some government programs offer tax incentives for
energy saving features. Other programs offer reduced
mortgage rates. The EnergyStar program offers simple
computer programs to determine the utility savings
and lease upgrades associated with energy-saving upgrades. Sustainable buildings often have features that
have been shown to increase worker productivity, decrease absenteeism, and increase student test scores
in schools.
The planned design of an energy-conserving or sustainable building requires the architect’s understanding of the effects of design decisions on energy performance. More than half of the true total costs incurred
during the economic life of a building may be attributable to operating and energy costs. An integrated
design approach considers how the walls interact with
the building and its HVAC system. Using this approach
early in the design phase helps optimize initial building costs and reduce long-term heating and cooling
energy costs.
Precast concrete panels have many inherent advantages when it comes to saving energy and protecting
1 ANSI/ASHRAE/IESNA Standard 90.1-1999 – Energy Standard for Buildings
Except Low-Rise Residential Building, American Society of Heating,
Ventilating, and Air-Conditioning Engineers (ASHRAE), Atlanta. http://
www.ashrae.org
2 Leadership in Energy and Environmental Design, U.S. Green Building
Council. www.USGBC.org
3 U.S. Environmental Protection Agency. www.EnergyStar.gov

392

|

ARCHITECTURAL PRECAST CONCRETE

the building from the environment. Their versatility
leads to unique solutions for energy conservation. The
relative importance of particular design strategies for
any given building depends to a large extent on its
location. For instance, buildings in northern, heatingseason-dominated climates are designed differently
than those in southern, cooling-season-dominated
climates.
Several factors influence the actual energy performance of the building envelope. Some of these are
recognized in energy codes and sustainability programs because they are relatively easy to quantify.
Others are more complex and are left to the discretion
of the designer.
Much of the information and design criteria that
follow are taken from or derived from the ASHRAE
Handbook of Fundamentals,4 and the ANSI/ASHRAE/
IESNA Standard 90.1. It is important to note that all
design criteria are not given and the criteria used may
change from time to time as the ASHRAE Handbook
and Standard are revised. It is therefore essential to
consult the applicable codes and revised references for
the specific values and procedures that govern in a particular area when designing the energy conservation
systems of a particular structure.
Building orientation plays an important role in
building energy consumption. If possible, the long axis
of the building should be oriented in the east-west direction to help control the effect of the sun on heating and cooling loads. If the long axis is parallel to the
east-west direction, solar gain through glazing on the
east side of the building in the morning and on the
west side in the afternoon increases the heat gain in
the building. This increases the air-conditioning load
on a building and makes it more difficult to control the
building temperature in different areas of a building.
However, east glazing will help warm an office building in early morning hours after night-temperature
setbacks.
To maximize beneficial solar heating, glazing should
be located on the south wall to maximize exposure to
winter sunshine in cold climates. South-facing glass
should be shaded to minimize solar exposure in the
summer.
In the southern regions of the U.S., the primary emphasis is on cooling. Glass should be more predomi4 ASHRAE Handbook of Fundamentals - 2001, American Society of Heating,
Refrigerating, and Air-Conditioning Engineers, Atlanta, GA. www.ASHRAE.org

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.2 Energy Conservation

nant on the north side of buildings in these regions to
minimize heat gains from the sun.
Building shape influences energy performance in
two ways. First, it determines the surface area of the
building skin. The larger the skin area, the greater the
heat gain (summer) or loss (winter). Second, shape influences how much of the floor area can be illuminated
using natural light from the sun, called daylighting. “E”
and “H” shaped buildings provide maximum exposure
of occupants to operable windows, but also have the
added benefit of providing optimal daylighting.
Glazing (the clear portion of windows) in buildings
requires special consideration during the design stage.
The type, amount, and orientation of glazing will profoundly affect heating, cooling, and daylighting requirements, HVAC system selection, human comfort,
and environmental satisfaction. Today’s high-performance glazing comes in many forms: those with low
emissivity (low-E), those filled with inert gas to further
lower U-factors, and those that are spectrally selective.
Heat gain through a sunlit glass area is many times
greater than through an equal area of precast concrete
and its effect is usually felt almost immediately. Direct
solar gains also cause glare in the work space. Properly
designed shading devices can modify the thermal effects of windows to a great extent. Glazing with low
solar heat gain and high visible light transmittance
provide the most benefits in most climates. More information on glazing is available through the National
Fenestration Rating Council (NFRC) (www.nfrc.org) and
the chapter on fenestration in the ASHRAE Handbook
of Fundamentals.
Daylighting saves energy by using natural light from
the sun rather than artificial lighting for illumination.
Controlling the type and amount of glazing influences
the benefits of daylighting. The potential energy savings from daylighting is particularly significant in commercial buildings because of the large lighting requirements in these buildings. Lighting can account for
approximately one-third of the building energy costs.
Daylighting controls can be used to dim or turn off
lights along the building perimeter when daylighting is
prevalent. Daylighting is not as effective as direct sunshine for providing light; as it is controlled low-glare
sunshine moderated by shading. Daylighting should be
maximized through location and size of windows and
through use of glazing systems and shading devices
appropriate to building orientation and space use.

5

Color (albedo) of precast concrete panels can be used
to improve the energy conserving features of the walls.
Panels with high albedo (generally lighter in color) can
help reduce the urban heat island effect. Albedo, which
in this case is synonymous with solar reflectance, is the
ratio of the amount of solar radiation reflected from a
material surface to the amount that shines on the surface. Solar radiation includes the ultraviolet, as well as
the visible spectrum. Albedo is measured on a scale of
0.0 to 1.0, from not reflective to 100% reflective, respectively. Generally, materials that appear to be lightcolored in the visible spectrum have high albedo and
those that appear dark-colored have low albedo, Table
5.3.1. Because reflectance in the solar radiation spectrum determines albedo, color in the visible spectrum
is not always a true indicator of albedo.
On exterior surfaces, high albedo surfaces (generally light colors) decrease solar heat gain; low albedo
(dark colors) increase solar heat gain. For instance, a
low-albedo north wall and high-albedo east and west
walls and roof form the most energy-conserving arrangement in a northern hemisphere climate that uses
both heating and cooling. For example, changing an
uninsulated building wall in Miami from a low albedo
to a high albedo can reduce annual cooling energy flux
(heat flow through the building envelope) by about
15%. High albedo surfaces are especially important
where cooling dominates the energy requirements. It
should be noted, however, that the color of the exterior walls has less effect on energy consumption when
the walls have high R-values and thermal mass. The
benefit of high-albedo surfaces in decreasing cooling
loads is often greater than the benefit of low-albedo
surfaces in decreasing heating loads even in cold climates. This occurs due to the decreased benefit of the
sun in the winter due to its lower angle, shorter days,
and often more cloudy conditions.
Light-colored exterior surfaces also help reduce urban
heat islands. Urban areas are up to 7°F warmer than
the surrounding areas, as shown in Fig. 5.3.1. This difference is attributed to more buildings and pavements
that have taken the place of vegetation. Trees provide shade that reduces temperatures at the surface.
Vegetation including trees provides transpiration and
evaporation that cool their surfaces and the air surrounding them. Where buildings and paved surfaces
are required, using materials with higher albedos will
reduce the heat island effect, save energy by reducing
the demand for air conditioning, and improve air qual-

ARCHITECTURAL PRECAST CONCRETE

|

393

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation

ity. The probability of smog greatly increases whenever
air temperatures exceed 75°F. Using trees and lightcolored surfaces can help reduce the number of hours
a city temperature is above 75°F, and thereby reduce
smog.

Fig. 5.3.1 Urban heat-island profile (LBNL website).

Planting deciduous trees that lose their leaves in the
winter, such as oak and maple, helps keep a building and the surrounding area cool. During the winter
months when no leaves are present, the building benefits from solar gains. Trees planted on the south and
west sides of building are particularly effective in providing shading and reducing solar gains in buildings.
Wind can decrease the exterior still-air film that usually surrounds a building and contributes to the insulating R-values of wall elements, thus increasing heating
and cooling loads. This effect is most predominant in
uninsulated concrete walls and becomes less marked
Table 5.3.1. Solar Reflectance (Albedo) of Select Material
Surfaces. 1, 2, 3, 4

Material Surface

Solar
Reflectance

Black acrylic paint

0.05

New asphalt

0.05

Black rubber or bitumen roof material

0.06

Aged asphalt

0.1

“White” asphalt shingle

0.2

Aged concrete

0.2 to 0.3

New concrete (traditional)

0.4 to 0.5

New concrete with white portland
cement

0.7 to 0.8

Aged average white membrane roof

0.77

White acrylic paint

0.8

Average white membrane roof

0.82

1 Levinson, Ronnen and Akbari, Hashem, “Effects of Composition and
Exposure on the Solar Reflectance of Portland Cement Concrete,”
Publication No. LBNL-48334, Lawrence Berkeley National Laboratory, 2001,
39 pp. http://eetd.lbl.gov/HeatIsland/
2 Pomerantz, M., Pon, B., and Akbari, H., “The Effect of Pavements’
Temperatures on Air Temperatures in Large Cities,” Publication No. LBNL43442, 2000, Lawrence Berkeley National Laboratory, 20 pp. http://eetd.
lbl.gov/HeatIsland/
3 Berdahl, P. and Bretz, S, “Spectral Solar Reflectance of Various Roof
Materials”, Cool Building and Paving Materials Workshop, Gaithersburg,
Maryland, July 1994 14 pp. http://eetd.lbl.gov/HeatIsland/
4 Pomerantz, M., Akbari, H., et al, “Examples of Cooler Reflective Streets for
Urban Heat Islands: Cement Concrete and Chip Seals,” Lawrence Berkeley
National Laboratory. http://eetd.lbl.gov/HeatIsland/

394

|

ARCHITECTURAL PRECAST CONCRETE

as the wall R-value and thermal mass increase. Wind
also carries solar heat away from a building and evaporates moisture on wet surfaces, thus possibly cooling
the skin to temperatures lower than the ambient air.
High winds create pressure differences across walls
which will increase air leakage through the walls. Cold
air leakage to the inside must be heated and probably
humidified. This also requires an expenditure of energy. Planting non-deciduous evergreen trees on the
windward (generally north and west) side of buildings
decreases energy losses in winter.
Texture of precast concrete panels has a minor effect on energy conservation. Increasing the surface
roughness of the wall exterior causes an increase in
the amount of sunlight absorbed and reduces the effect of wind on heat loss and gain. Ribbed panels act
as baffles to wind, thereby reducing conductive heat
loss and infiltration. Although this has a somewhat
smaller effect than proper color selection, it can help
to reduce total energy consumption. However roughness and ribs can also decrease solar reflectivity and
increase solar heating, thus indicating a balance is necessary depending on local conditions.
Air infiltration has significant effects on the amount
of energy required to heat and cool a building. Air leaks
into or out of the building envelope through gaps between building materials. The amount of leakage is dependent on the size of the gaps and pressure differences due to building height, indoor-outdoor temperature
differences, and wind pressure. Air leakage increases
as pressure differences increase. Additional information
on air infiltration is provided in section 5.3.6.
Shading is a fundamental design strategy for pre-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.2 Energy Conservation

Shading using horizontal or vertical plane(s) projecting
out in front of or above a window can be designed to
block the summer sun, allow most of the winter sun,
and provide a view for occupants. If the plane projects
far enough from the building, a single projection may
be sufficient, as in the case of generous roof overhangs
or windows recessed deeply between vertical fins.
Alternatively, more modest projections can be equally effective in shading but they must be more closely
spaced. Closely spaced horizontal or vertical planes may
begin to dominate the view out of a window and in any
case change the scale of the window. The proportion
of the space divided by the shading planes becomes as
important as the overall window proportion in determining the aesthetic effect of the fenestration.
In summer, vertical fins will shade the early morning
and late afternoon sun while horizontal fins keep out
the high-altitude mid-day sun. In winter these shades
will not interfere with the sun because of its low altitude and southerly azimuth at sunrise and sunset.
Horizontal shading is most effective on southern exposures, but if not extended far enough beyond the
windows, it will permit solar impingement at certain
times of the day. Designs may be flat or sloping; slop-

Fig. 5.3.2 Daylighting – panels used to shade glass.
Summer Sun

Winter Sun
Multiple Fins
May Be Cast
Integrally

Glass

Wint
er So
lstice
Sola
r Ang
le 25
°

venting solar heat gain and diffusing bright sunlight.
Recessed window walls, vertical fins, and various other
sculptured shapes facilitate the design of many types
of shading devices for windows, including vertical and
horizontal sunshades. In the cooler months, when the
sun’s angle of incidence is low, the shading devices may
be angled to let the sunshine in and help reduce heating loads on the building’s heating system, (Fig. 5.3.2).
The shading approach selected can reinforce and enhance the design content and form of the building, in
some cases becoming the prime form-giving element.
Shading may have to be modified or compromised in
order to meet other important requirements. Figure
5.3.2 shows preferred cross-sections (in elevation) for
precast concrete used as shading elements. Note that in
each case, the spandrel and sunscreening elements are
integral and may be lifted into place in one operation.
The designer should be aware of the increased possibility of glass breakage from sharp shading lines caused
by a thermal stress differential between the shaded and
unshaded portions of a singleglass unit. Where shading
differential is anticipated, the use of heat-strengthend
rather than annealed glass is often advised by glass
manufacturers.

5

ice
olst
er S
m
70°
Sum
ngle
A
r
Sola

May Be Welded
To Panel

May Be Welded
To Panel

ing versions may be of shorter length, but obstruct
more of the sky view (Figs. 5.3.3 and 5.3.4). The detached screen panel parallel to the wall in Fig. 5.3.5
was used to block the rays of the sun, while still allowing light to enter the windows. Sun-shading may
also be provided through the use of a free-standing
perimeter structure set in front of the actual building
enclosure (Fig. 5.3.6).

ARCHITECTURAL PRECAST CONCRETE

|

395

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation

Horizontal sunscreen.

Fig. 5.3.3
Cornerstone, Plantation, Florida; Architect: Thompson, Ventulett, Stainback & Associates; Photos: Brian Gassel/TVS.

Horizontal shading can have a significant impact
on heat gain through windows. In Miami, overhangs
with a projections factor of 0.5 can reduce annual energy flux (heat flow through the building envelope) by
about 15%. The relative impact declines to about a
10% reduction in northern climates. A projection factor is defined as the horizontal length of the overhang
divided by distance from the bottom of the glass to the
underside of the projection. So a projection about half
the height of the window, directly above the window,
will have a projection factor of about 0.5. Permanent
projections can be used to help meet the solar heat

396

|

ARCHITECTURAL PRECAST CONCRETE

gain coefficient (SHGC) requirement when using ANSI/
ASHRAE/IESNA Standard 90.1-2007.
In windy areas, the solar screens can be made to serve
the double purpose of wind-breaks. Trees adjacent to
the building can also serve the function of sun shading
and wind-breaks.
Sunscreen panels, which have pockets to receive precast double tees, form the south, east and west faces
of the midrise office building in Fig. 5.3.7(a) and (b)
while the north face features flat panels with punched
openings.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.2 Energy Conservation

Horizontal sunscreen.

Fig. 5.3.4
Blue Cross and Blue Shield of Connecticut, North Haven,
Connecticut; Architect: Ellenzweig Associates,Inc.;
Photo: ©1990 Steve Rosenthal.

5

Detached screens.

Fig. 5.3.5
United Parcel Service Corporate Offices, Atlanta, Georgia;
Architect: Thompson, Ventulett, Stainback & Associates;
Photo: Brian Gassel/TVS.

Fig. 5.3.6
Florida Atlantic University Library, Boca Raton, Florida;
Architect: Spillis Candela DMJM formerly Spillis Candela &
Partners, Inc.; Photo: Spillis Candela DMJM.

Free standing screen.

ARCHITECTURAL PRECAST CONCRETE

|

397

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation

(b)

Fig. 5.3.7(a) & (b)
Denver Corporate Center,
Denver, Colorado;
Architect: Hammond Beeby and Babka, Inc.;
Photos: Balthazar Korab Ltd.

Variation of panel shape on different faces.

(a)

Variation of window openng orientation.

Fig. 5.3.8
Valley River Office Park, Eugene,
Oregon; Architect: Boutwell,
Gordon, Beard and Grimes;
Photo: Boutwell, Gordon, Beard
and Grimes.

398

|

ARCHITECTURAL PRECAST CONCRETE

Solar control through the use of shading devices is
most effective when designed specifically for each façade, since time and duration of solar radiation vary
with the sun’s altitude and azimuth. The designer can
predict accurately the location and angles of the sun,
designing overhangs or fins to shade exactly the area
desired. This type of envelope response can be seasonal (shade during certain times of the year) or daily
(shade during certain hours of the day).
The versatility of precast concrete was used to change
the window opening configuration with respect to
wall orientation in order to maximize solar gains in the
winter and minimize them in the summer (Fig. 5.3.8).
Since the windows are small relative to the wall surface, the window units were splayed back on two different planes (at the sill and jamb) so that the windows
could be recessed and shaded.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.2 Energy Conservation

East and west facing windows are more effectively shaded by vertical projecting planes (Fig. 5.3.9).
Vertical projections (fins) from either side of the window narrow the peripheral view from the window. In
the Northern hemisphere the further south a building
is located, the more important shading east- and westfacing windows becomes, and the less important it is
to shade south-facing windows. This is due to the high

5

position of the summer sun in southern latitude with
the resulting decrease in direct sunlight transmitted by
the south-facing windows.
In Fig. 5.3.10, the top floor is cantilevered over the
main floor to shade the windows. All second floor windows on the east and west sides are oriented directly
south or north for sun control. The vertical wingwall

Use of vertical fins.

Fig. 5.3.9
Medical Science Research Building, Duke University Medical Center, Durham, North Carolina; Architect: Payette Associates;
Photo: ©2007 Brian Vanden Brink, Photographer.

ARCHITECTURAL PRECAST CONCRETE

|

399

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation

shading devices completely shade the windows during
the four summer months.
The use of three-dimensionally profiled precast concrete window wall units permits windows to be recessed
within an enclosing concrete surround. The sides may
be vertical or angled. Deeply recessed windows are particularly effective in minimizing solar heat gains without reducing natural light and view. Eggcrate shading
works well on walls facing southeast, and is particularly effective for a southwest orientation. Because of
its high shading efficiency, the eggcrate device (deeply
recessed windows) is often used in hot climates. The
deep, recessed window areas and massive overhangs
in Fig 5.3.11 illustrate the total flexibility of design that
precast concrete offers the architect.

Three-foot-deep (1 m) “eyebrows,” were the shading
device used to keep out the sun’s rays in the summer
and reduce cooling loads (Fig. 5.3.12).
Precast concrete and inclined glass can work together
for optimum use of daylighting. Direct sun strikes the
glass at an angle and is reflected, reducing glare, while
indirect sunlight reflects off the sill of the precast concrete panel and through the glass to provide soft natural light at the perimeter of the building (Fig. 5.3.13). By
keeping the direct rays of the sun out of the building,
cooling loads are considerably reduced and daylighting is maximized. “Eyelid” or hooded shading devices
and inclined glass can be very effective in controlling
the penetration of the sun into a building by reducing the area of glass exposed to the sun (Fig. 5.3.14).

Cantilevered floor used to shade windows.

Fig. 5.3.10
Arizona Public Service Administration Complex, Phoenix, Arizona; Architect: DFD Cornoyer-Hedrick formerly Comoyer-Hedrick Architects &
Planners; Photo: DFD Comoyer Hedrick.

400

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.2 Energy Conservation

5

Deep recessed windows.

Fig. 5.3.11
East Los Angeles Municipal Courts Building, East Los Angeles, California; Architect: Kanner Architects; Photo: Kanner Architects.

ARCHITECTURAL PRECAST CONCRETE

|

401

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.2 Energy Conservation / 5.3.3 Thermal Resistance (R-value)

Deep recessed windows and overhang (eggcrate device).

Fig. 5.3.12
Strom Thurmond Federal Office Building and Courthouse, Columbia, South Carolina; Architect: Marcel Breuer Associates and Design
Partnership; Photo: ©Image courtesy of Marcel Breuer papers, 1920-1986, in the Archives of American Art, Smithsonian Institution.

5.3.3 Thermal Resistance (R-value)
This shading device softens the brightness contrast between the interior and exterior. Rounded heads, sills,
and jambs or deep window wells could also be used to
soften brightness contrasts (Fig. 5.3.15).

402

|

ARCHITECTURAL PRECAST CONCRETE

Common thermal properties of materials and air spaces are based on steady state tests, which measure the
heat that passes from the warm side to the cool side of
the test specimen. Thermal mass of concrete, which is

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.3 Thermal Resistance (R-value)

5

Inclined glass.

Fig. 5.3.13
Security Insurance Group Headquarters, Farmington, Connecticut; Architect: Russell Gibson von Dohlen; Photo: Steve Rosenthal.
Hooded shading device.

Fig. 5.3.14
Miami Police Station, Miami, Florida; Architect: Borrelli
+ Partners formerly Pancoast, Bouterse, Borrelli, Albaisa,
Architects/Planners, Inc.; Photo: Borrelli + Partners.

Fig. 5.3.15 Jambs, head or sill (rotate 90º).

3TRAIGHT
Straight

2OUNDED
Rounded

!NGLED
Angled

ARCHITECTURAL PRECAST CONCRETE

|

403

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.3 Thermal Resistance (R-value)

not based on steady-state tests, is discussed in Section
5.3.5. Daily temperature swings and heat storage effects are accounted for in thermal mass calculations.
The results of steady-state tests provide the thermal
resistance (R-value) of the air, material, or combination
of materials tested. Tests of homogenous materials are
also used sometimes to provide the thermal conductivity value. The R-value per inch of a homogenous material is equal to the inverse of its thermal conductivity.
The R-value for a material with a specific thickness is its
thickness divided by thermal conductivity.

als (Rmaterials) in the section, the indoor and outdoor air
film surfaces (Rfi and Rfo) and air spaces (Ra) within the
section.
Rtotal = Rfi + Rmaterials + Ra + Rfo

Equation 5.3.1

or
Rtotal = Rfi + Rconcrete + Rinsulation + Ra + Rfo
These equations are only applicable for layered systems where each layer is composed of a homogenous
material. In framing or other systems where members
or elements penetrate the insulation layer, the seriesparallel or zone method from the ASHRAE Handbook
of Fundamentals must be used.

The overall (total) R-value of a building wall is computed by adding together the R-values of the materiTable 5.3.2 Thermal Resistances, Rf, of Surfaces1.

Outdoor – Moving Air, Rfo

Indoor – Still Air, Rfi
Reflective surface

Non-reflective surface

Position of
surface

Direction of
heat flow

Nonreflective
surface

Aluminumcoated paper,
polished

Bright
aluminum
foil

Vertical

Horizontal

0.68

1.35

1.70

0.17

0.25

Up

0.61

1.10

1.32

0.17

0.25

Down

0.92

2.70

4.55

0.17

0.25

Horizontal

15 mph wind, 7.5 mph wind,
winter design summer design

1. ASHRAE Handbook of Fundamentals, 2005, www.ASHRAE.org.

Table 5.3.3 Thermal Resistances, Ra, of Air Spaces1.

Air Space
Position of
Air Space

Direction of
Heat Flow
Horizontal
(walls)

Vertical
Horizontal
(walls)

Mean
Temp.
temp., °F difference, °F
Winter
50
10
50
30
Summer
90
10

NonReflective
Surfaces

Reflective Surfaces
One side2 One side3 Both sides3

1.01
0.91

2.32
1.89

3.40
2.55

3.63
2.67

0.85

2.15

3.40

3.69

10
30
30
10

0.93
0.84
1.22
1.24

1.95
1.58
3.86
4.09

2.66
2.01
8.17
9.27

2.80
2.09
9.60
11.15

10

1.00

3.41

8.19

10.07

Winter
Up (roofs)
Horizontal

Down (floors)
Down (roofs)

50
50
50
50
Summer
90

1 For 31/2 in. air space thickness. The values, with the exception of those for reflective surfaces, heat flow down, will differ about 10% for air space thicknesses
of 3/4 in. to 6 in. Refer to the ASHRAE Handbook of Fundamentals for values of other thicknesses, reflective surfaces, heat flow directions, mean temperatures,
and temperature differentials. ASHRAE Handbook of Fundamentals, 2005, www.ASHRAE.org.
2 Aluminum painted paper.
3 Bright aluminum foil.

404

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.3 Thermal Resistance (R-value)

The U-factor is the reciprocal of the total R-value
(U = 1/Rtotal).
Tables 5.3.2 and 5.3.3 give the thermal resistances of
air films and 31/2 in. (90 mm) air spaces, respectively.
The R-values of air films adjacent to surfaces and air
spaces differ depending on whether they are vertical,
sloping, or horizontal and, if horizontal, whether heat
flow is up or down. Also, the R-values of air films are
affected by the velocity of air at the surfaces and by
their reflective properties.
Tables 5.3.4 and 5.3.5 provide thermal
properties of most commonly used building
materials. The R-values of most construction materials vary somewhat depending on
the temperature and thickness. Note that
expanded polystyrene and extruded polystyrene board insulation have different thermal
and physical properties. Expanded polystyrene (EPS) or beadboard is composed of small
beads of insulation fused together. Extruded
polystyrene (XPS) is usually pigmented blue,
pink, or green, and has a continuous closed
cell structure. XPS generally has a higher thermal resistance, higher compressive strength,
and reduced moisture absorption compared
to EPS. Mineral fiber and fiberglass batt insulation are not included in the table, but are
generally labeled by the manufacturer. The
most common batts for walls are R11, R13,
and R19, with the number following the R indicating the R-value.
Glazing thermal performance is measured by
thermal transmittance (U-factor), solar heat
gain coefficient (SHGC), and visible light transmission (VLT). A low SHGC will minimize solar
heat gains and reduce cooling loads. Some
products with low SHGC also have a low VLT
that will reduce daylighting benefits. Products
with a low SHGC and high VLT are often a
good choice. Since glazing types have proliferated in recent years, refer to the ASHRAE
Handbook of Fundamentals or the NFRC for
glazing and fenestration properties. Table
5.3.5 provides some typical values.
Table 5.3.6 gives the thermal properties of
various weight concretes in the “normally dry”
condition. Normally dry is the condition of concrete containing an equilibrium amount of free

5

water after extended exposure to room temperature air
at 35 to 50 percent relative humidity. Thermal conductivities and resistances of other building materials are
usually reported for oven dry conditions. However, concrete starts out wet and is rarely in the oven dry condition. Higher moisture content in concrete causes higher
thermal conductivity and lower thermal resistance.
However, normally dry concrete in combination with
insulation generally provides about the same R-value as
equally insulated oven dry concrete.

Table 5.3.4 Thermal Properties of Various Building Materials at 75°F 1.

Density,
lb/ft3

Resistance,
R per in. of
Thickness,
hr·ft2·°F/Btu

Specific
Heat,
Btu/(lb·°F)

8.0
4.0 – 9.0
15

3.03
3.1 – 4.2
3.1 – 4.2

0.18
0.23
0.17

Extruded polystyrene (XPS),
extruded cont. closed cell

1.8 – 3.5

5.00

0.29

1.0

3.85

—

Expanded polystyrene (EPS),
molded bead

1.25
1.5
1.75

4.00
4.17
4.17

2.0

4.35

—

Cellular polyurethane/
polyisocyanurate (unfaced)

1.5

6.25 – 5.562

0.38

Cellular phenolic, closed cell
Cellular phenolic, open cell
Polyicynene

3
1.8 – 2.2
0.5

8.2
4.4
3.6

50
50

0.88
1.06

0.26
0.31

116

0.20

0.20

45
105
38 – 47
24 – 41
34

0.63
0.18
0.94 – 0.80
1.35 – 0.89
1.25

—
0.20
0.39
0.39
0.29

Material
Insulation, rigid
Cellular glass
Glass fiber, organic bonded
Mineral fiber, resin bonded

Miscellaneous
Gypsum board
Particle board
Plaster
cement, sand aggregate
gypsum, lightweight
aggregate
gypsum, sand aggregate
Wood, hard (maple, oak)
Wood, soft (pine, fir)
Plywood

1 See Table 5.3.6 for concrete. ASHRAE Handbook of Fundamentals, www.ASHRAE.org.
2 An aged value of 6.0 is currently recommended. Environmental Building News, January 2005,
www.BuildingGreen.com.

ARCHITECTURAL PRECAST CONCRETE

|

405

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.3 Thermal Resistance (R-value)

Table 5.3.5 Thermal Properties of Window with an Aluminum Frame1.

U-Factor,
Btu/hr·ft2·ºF SHGC2

Window System
Double glazing with low-E coating and argon gas fill in an aluminum frame with thermal break

0.36

0.36

Double glazing with a low-E coating in an aluminum frame with thermal break

0.40

0.36

Double glazing in an aluminum frame with thermal break

0.56

0.65

Single glazing in an aluminum frame with no thermal break

1.13

0.65

1 Values will vary by manufacturer; check with supplier.
2.SHGC can vary significantly up or downward with the coating selected.

Table 5.3.6 Thermal Properties of Concrete1.

Resistance, R
Description

Concrete
including
normal weight,
lightweight,
and lightweight
insulating
concretes

Normal weight
solid panels,
140 to 150 pcf,
sand and gravel
aggregate

Structural
lightweight
solid panels

Concrete Density, Thickness,
lb/ft3
in.
145
140
130
120
110
100
90
80
70
60
50
40
30
20

145

110

Per in. of thickness, For thickness shown, Specific heat,
hr·ft2·°F/Btu
hr·ft2·°F/Btu
Btu/(lb·°F)
0.063
0.20
0.068
0.20
0.083
0.20
0.10
0.20
0.13
0.20
0.16
0.20
0.21
0.20
0.27
0.20
0.36
0.20
0.44
0.20
0.59
0.20
0.71
0.20
0.91
0.20
1.25
0.20

2
3
4
5
6
8

0.13
0.19
0.25
0.31
0.38
0.50

2
3
4
5
6
8

0.26
0.38
0.51
0.64
0.76
1.02

0.20

0.20

1 Based on values in the 2005 ASHRAE Handbook of Fundamentals and ANSI/ASHRAE/IESNA Standard 90.1-2007. Values do not include air film resistances. See Table 5.3.7
for R-values with air film resistances.

406

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.3 Thermal Resistance (R-value)

5

Table 5.3.7 R-Values for Solid Concrete and Sandwich Panel Walls1.

Concrete
Density,
lb/ft3

R-Value
of
Concrete
t,2 in.
No Air
Films, No
Insulation
2
3
4
5
6
8
2
3
4
5
6
8

145

110

0.13
0.19
0.25
0.31
0.38
0.50
0.25
0.38
0.51
0.64
0.76
1.02

R-Value of Insulation Resistance, hr·ft2·ºF/Btu
0
(No
ins.)

1

2

3

4

5

6

8

10

12

15

16

18

0.98
1.04
1.10
1.16
1.23
1.35
1.10
1.23
1.36
1.49
1.61
1.87

2.0
2.0
2.1
2.2
2.2
2.4
2.1
2.2
2.4
2.5
2.6
2.9

3.0
3.0
3.1
3.2
3.2
3.4
3.1
3.2
3.4
3.5
3.6
3.9

4.0
4.0
4.1
4.2
4.2
4.4
4.1
4.2
4.4
4.5
4.6
4.9

5.0
5.0
5.1
5.2
5.2
5.4
5.1
5.2
5.4
5.5
5.6
5.9

6.0
6.0
6.1
6.2
6.2
6.4
6.1
6.2
6.4
6.5
6.6
6.9

7.0
7.0
7.1
7.2
7.2
7.4
7.1
7.2
7.4
7.5
7.6
7.9

9.0
9.0
9.1
9.2
9.2
9.4
9.1
9.2
9.4
9.5
9.6
9.9

11.0
11.0
11.1
11.2
11.2
11.4
11.1
11.2
11.4
11.5
11.6
11.9

13.0
13.0
13.1
13.2
13.2
13.4
13.1
13.2
13.4
13.5
13.6
13.9

16.0
16.0
16.1
16.2
16.2
16.4
16.1
16.2
16.4
16.5
16.6
16.9

17.0
17.0
17.1
17.2
17.2
17.4
17.1
17.2
17.4
17.5
17.6
17.9

19.0
19.0
19.1
19.2
19.2
19.4
19.1
19.2
19.4
19.5
19.6
19.9

1 Values in table are the total R-values of the walls with concrete of thickness t and insulation R-value as indicated in columns. Only for insulation with no metal or
solid concrete penetrating the insulation layer. R-values will be impacted by the presence of these items and additional calculations will be required according to
series-parallel or zone method. Air film resistances of 0.68 for inside and 0.17 for outside are included in R-values unless otherwise noted. These are standard air
film resistances for winter conditions and are conservative for summer conditions.
2 The thickness, t, is the sum of the thicknesses of the concrete wythes for a sandwich panel wall.

A number of typical concrete wall R-values are given
in Tables 5.3.6 and 5.3.7. These wall tables can be applied to sandwich type panels, as well as single wythe
panels insulated on one side. The U-factor of the wall
is the inverse of the R-value with air film resistances
from Table 5.3.7. To use Table 5.3.7, first determine
the R-value of the insulation to be used either from
Table 5.3.4 or from the insulation manufacturer.
Manufacturers of insulation are required by law to provide the R-value of their material.

The R-value of walls assemblies are generally only calculated for the winter condition since the difference
between the summer and winter conditions is small.
This example is valid only for insulation with no metal
or solid concrete penetrating the insulation layer. Rvalues will be impacted by the presence of these items
and additional calculations will be required according
to series-parallel, zone, or characteristic method.

For concrete walls with metal furring or studs, wall Rvalues can be determined using Tables 5.3.7 and 5.3.8.
Determine the R-value of the concrete portion from
Table 5.3.7 and add it to the effective R-value from the
insulation/framing layer from Table 5.3.8, page 411.

Example 5.3.1 – Calculate R-Value of Wall Assembly.

C
A

D
B

E

The following design
example shows how to
calculate R-value and Ufactor for a wall using
material R-values taken
from Tables 5.3.2 through
5.3.7.

Thermal bridges such as metal wythe connectors

Wall Layer
A. Surface, outside air film
B. Concrete, 2 in. (145 pcf)

R
R
Table
Winter Summer
0.17
0.25
5.3.2
0.13
0.13
5.3.7

C. EPS insulation
(1.25 pcf), 1.5 in.

6.00

6.00

5.3.4

D. Concrete, 2 in. (145 pcf)
E. Surface, inside air film
Total R =
U = 1/R

0.13
0.68
7.11
0.14

0.13
0.68
7.19
0.14

5.3.7
5.3.2

ARCHITECTURAL PRECAST CONCRETE

|

407

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.3 Thermal Resistance (R-value)

or a full thickness of concrete along sandwich panel
edges will reduce the R-value of the wall. The net effect of metal ties is to increase the U-value by 10 to
15 percent, depending on type, size, and spacing. For
example, a wall as shown in Fig. 5.3.16 would have a
U-value of 0.13 if the effect of the ties is neglected.
If the effect of 1/4 in. diameter ties at 16 in. on center is included, U = 0.16; at 24 in. spacing, U = 0.15.
Ongoing research indicates these numbers are conservative. As another example, steel ties representing
0.06 percent of an insulated panel area can reduce the
panel R-value by 7 percent.5
Thermal bridging is minimized by the use of engineered resin, low conductivity wythe connectors in insulated concrete panel construction. These composite
material connectors, along with their ability to enable
edge to edge insulation coverage in the concrete sandwich panels, can significantly reduce thermal bridging
and help the insulation layer to retain up to 99.7 percent of its listed R-value.
Thermal bridges may lead to localized cold areas
where surface condensation can occur, particularly
where the interior relative humidity is maintained at
high levels. This may cause annoying or damaging wet
streaks on the wall surface. Icicles have been reported
on the interior side of some buildings in cold climates,
see Section 5.3.6. In most cases the problem has been
traced to excessive air exfiltration through major openings in the wall, often at precast concrete wythe connector locations. Since steel connectors form a high
conductivity path, they offer likely locations for condensation to occur. Corrosion protection, stainless
steel, or increased thickness of the connector material
may provide extended service life for these steel wythe
connectors.
The effect of metal tie thermal bridges on the heat
transmittance may be calculated by the zone method
described in the ASHRAE Handbook of Fundamentals
although the characteristic method is preferred. With
the zone method, the panel is divided into Zone A,
which contains the thermal bridge, and Zone B, where
thermal bridges do not occur, as shown in Figure
5.3.16. The width of Zone A is calculated as W = m +
2d, where m is the width or diameter of the metal or
other conductive bridge material, and d is the distance
5 VanGeem, M. G., "Effects of Ties on Heat Transfer Through Insulated
Concrete Sandwich Panel Walls," Proceedings of the ASHRAE/DOE/BTECC/
CIBSE Conference on Thermal Performance of the Exterior Envelopes of
Buildings IV, Orlando, December 1989, ASHRAE, Atlanta, 1989, pp. 206223. www.ASHRAE.org

408

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.16 Metal tie thermal bridges.

8"
3" 2" 3"

W

Welded wire
reinforcement
Outside

Inside
Zone B

Ties
Rigid
Insulation

Zone A

from the panel surface to the metal. After the width
(W) and area (A) of Zone A are calculated, the heat
transmissions of the zonal sections are determined and
converted to area resistances, which are then added
to obtain the total resistance (Rt) of that portion of the
panel. The resistance of Zone A is combined with that
of Zone B to obtain the overall resistance and the gross
transmission value Uo, where Uo is the overall weighted
average heat transmission coefficient of the panel.
The effect of solid concrete path thermal bridges can
be calculated by the characteristic section method.
In this method, the panel is divided into two regions.
The first region is treated as a perfectly insulated panel
without any thermal bridge. The second region is treated as a solid concrete panel without any insulation.
The total thermal resistance of the panel is calculated
as the resistances of these two regions added together
in parallel.
The portion of the panel that is treated as a solid concrete panel without any insulation is larger than the
actual solid concrete region that exists in the panel.
There is an affected zone around each solid concrete
region that is added to the actual area of the solid concrete region to obtain the size of the concrete region
used in the calculation. The size of the affected zone
Ez is computed as:
Ez = 1.4 - 0.1tinα + 0.4t cf + 0.1( t cb - t cf )  β
Equation 5.3.2

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.3 Thermal Resistance (R-value)

1 ft.

12 in. + 2 x (2.3) in. = 16.6 in.
12 in. + 2 x (2.3) in. = 16.6 in.

Equation 5.3.4

1 ft.

Transverse Section

In these equations, kin and kcon have units of Btu
(in/hr)(ft2)(°F).
To calculate an R-value, a panel is divided into two regions: a solid concrete region and a perfectly insulated
region, as explained previously. Ez is calculated using
Equation 5.3.2 and the area of each region is then calculated. The thermal resistance of the solid concrete
region (Rs) is then added in parallel with the thermal
resistance of the perfectly insulated region (Rp) to obtain the thermal resistance of the panel R:
1 A ′s Ap′
=
+
R Rs Rp

12 in. + 2.3 in. = 14.3 in.
1 ft.

Equation 5.3.3

and
 k - 12.05 
β = 1+ 1.458  con
 12.05 

12 ft.

40 ft.

 k - 0.26 
α = 1+ 2.25  in
 0.26 

Example 5.3.2 Determination of R-value for sandwich panel.

1 ft.

In this equation, tin, tcf and tcb are the thicknesses of
the insulation layer, concrete face wythe, and concrete
back wythe, respectively. This is an empirical equation
with all dimensions expressed in inches. The parameters α and β account for the insulation and concrete
conductivity values (kin and kcon) that are used to construct the panel. Their values are computed as:

5

Equation 5.3.5

A′s and A′p represent the areas of the solid concrete
region (As) and perfectly insulated panel region (A p) divided by the total panel area At (i.e. A′s = A s /A t, A′p =
A p /At). The procedure is illustrated in Example 5.3.2.
Where:

Vertical Section

Affected Zones

tin

= thickness of insulation layer

α

= insulation conductivity coefficient factor

β

= concrete conductivity coefficient factor

Problem:
Determine the R-value for the sandwich panel shown
above for conductivities of 10.0 Btu·(in./hr)·(ft2)°F and
0.15 Btu·(in./hr)·(ft2)(°F) for the concrete and insulation, respectively. Face and back wythe thicknesses are
3 in., and the insulation layer thickness is 2 in.
Solution:
Calculate the parameters α and β:
 k - 0.26 
 0.15 - 0.26 
= 1+ 2.25 
a = 1+ 2.25  in.
= 0.05
0.26 
 0.26 


Ap = area of insulated panel zone
As = area of solid concrete zone
At = total area of panel
A′ = portion of each zone
A′p = portion of insulated panel zone
A′s = portion of solid concrete zone
Ez

= affected zone

kcon = conductivity of concrete
kin = conductivity of insulation
tcb = thickness of back concrete wythe
tcf

= thickness of face concrete wythe

 k - 12.05 
 10.00 - 12.05 
b = 1+ 1.458  con
= 1+ 1.458 


12.05
 12.05 


= 0.75
From the panel thicknesses, the affected zone dimension Ez is computed as:
Ez = 1.4 – 0.1(tin)(α) + [0.4tcf + 0.1(tcb – tcf)] β
Ez = 1.4 – 0.1(2)(0.05) + 0.4(3)(0.75)
Ez = 2.3 in.
Add Ez to the actual solid concrete areas to obtain the
areas of the panel to treat as solid concrete (shown as
dashed lines above).

ARCHITECTURAL PRECAST CONCRETE

|

409

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.3 Thermal Resistance (R-value) / 5.3.4 Heat Capacity

Calculate the areas of the panel (At), solid concrete
region (As), and perfectly insulated region (Ap):
At = panel area = (40 ft)(12 ft) = 480 ft2 = 69,120
in.2
As	= concrete area = 2(14.3)(144) + 8(16.6)(16.6)
= 6,323 in.2
Ap = insulated area = 69,120 – 6,323 = 62,797 in.2
This resistance of that portion of the panel that is
treated as perfectly insulated is calculated from the resistances of the concrete, insulation, and surfaces in
series.
The resistance of that portion of the panel that is
treated as solid concrete is calculated from the resistances of the concrete and surfaces in series.
Calculate the fractional areas of the panel that are
treated as solid concrete and as insulated:
Insulated Path.

k
A

Outside surface

B

Thickness U = k/t

R = 1/U
Winter

—

—

—

0.17

Concrete

10.00

3

3.33

0.30

C

Insulation

0.20

2

0.10

10.00

D

Concrete

10.00

3

3.33

0.30

—

—

—

E

Inside surface
Total

0.68
11.45

As /At = 6323/69120 = 0.091
Ap /At = 62797/69120 = 0.909
Concrete Path.

k

Thickness U = k/t

R = 1/U
Winter

A

Outside surface

—

—

—

0.17

B

Concrete

10.00

8

1.25

0.80

C

Insulation

—

—

—

0.68

Total

1.65

Compute the R-value of the panel treating the solid
concrete and perfectly insulated regions in parallel.
1 0.909 0.091
=
+
R 11.45 1.65

410

|

1 0.909 0.091
=
+
R 11.53 1.73

ARCHITECTURAL PRECAST CONCRETE

Winter:

Summer:

R = 7.43 hr · ft2 · °F/Btu

R = 7.61 hr · ft2 · °F/Btu

ASHRAE Standard 90.1 also recognizes the detrimental thermal bridging effects of steel framing within
walls. For example, ASHRAE specifies an effective insulation/framing R-value of 5.1 for R13 insulation in a 4
in. metal stud cavity for concrete wall construction. For
the effects of other metal framing depths and insulation R-values in precast concrete walls see Table 5.3.8.

5.3.4 Heat Capacity
Heat Capacity (HC) is used in energy codes to determine when a wall has enough thermal mass to use the
mass criteria or mass credit. Heat capacity is the ability
to store heat per unit area of wall area and includes
all layers in a wall. For a single layer wall, HC is calculated by multiplying the density of the material by its
thickness times the specific heat of the material. Heat
capacity for a multilayered wall is the
sum of the heat capacities for each
R = 1/U
layer. The heat capacity of non-conSummer
crete layers is generally small and can
0.25
typically be ignored in calculations.
0.30
Specific heat describes a material’s
10.00
ability
to store heat energy. As a ma0.30
terial absorbs energy, its temperature
0.68
rises. A material with a high specific
11.53
heat, such as water, can absorb a
great deal of heat energy per pound
of material, with little rise in temperature. The same
weight of a material with low specific heat, such as
steel or copper, rises to higher temperatures with only
a small quantity of heat absorbed. Because specific
heat defines the relationship between heat energy and
temperature for a given weight of maR = 1/U
terial, it can also be used to determine
Summer
the change in temperature for a ma0.25
terial as it absorbs or releases energy.
0.80
Specific heat is defined as the quantity of heat energy in Btus required to
0.68
raise the temperature of one pound of
1.73
a material by 1ºF. The specific heat of
concrete can generally be assumed to be 0.2 Btu/lb·ºF.
Specific heat of selected other materials is provided in
Table 5.3.4.
Energy codes generally require a heat capacity greater than 6 Btu/ft2·°F in order to use mass

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.4 Heat Capacity

5

Depth of framing
and cavity, (in.)

Table 5.3.8 Effective R-Values for Walls with Insulation in Cavity between Metal Furring or Studs1.

Depth of framing
and cavity, (in.)

0.5
0.8
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5

0.5
0.8
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5

Rated R-value of insulation
0

1

2

3

4

5

6

7

8

9

10

11

12

Effective R-value if continuous insulation uninterrupted by framing (includes gypsum board)
0.5

1.5

2.5

3.5

4.5

5.5

6.5

7.5

8.5

9.5

10.5

11.5

12.5

Effective R-value if insulation is installed in cavity between metal framing (includes gypsum board)
0.9
1
1
1.1
1.1
1.2
1.2
1.2
1.2
1.2
1.2
1.3

0.9
1
1.1
1.2
1.2
1.3
1.3
1.3
1.3
1.3
1.4
1.4

1.1
1.3
1.4
1.6
1.7
1.8
1.9
2
2
2.1
2.1
2.1

1.1
1.4
1.6
1.9
2.1
2.3
2.4
2.5
2.6
2.6
2.7
2.8

1.2
1.5
1.7
2.1
2.3
2.6
2.8
2.9
3
3.1
3.2
3.3

na
1.5
1.8
2.2
2.5
2.8
3.1
3.2
3.4
3.5
3.7
3.8

na
1.6
1.8
2.3
2.7
3
3.3
3.5
3.7
3.9
4.1
4.2

na
na
1.9
2.4
2.8
3.2
3.5
3.8
4
4.2
4.4
4.6

na
na
1.9
2.5
2.9
3.3
3.7
4
4.2
4.5
4.7
4.9

na
na
na
2.5
3
3.5
3.8
4.2
4.5
4.7
5
5.2

na
na
na
2.6
3.1
3.6
4
4.3
4.6
4.9
5.2
5.4

na
na
na
2.6
3.2
3.6
4.1
4.5
4.8
5.1
5.4
5.7

na
na
na
2.7
3.2
3.7
4.2
4.6
5
5.3
5.6
5.9

17

18

19

20

21

22

23

24

25

Rated R-value of insulation
13

14

15

16

Effective R-value if continuous insulation uninterrupted by framing (includes gypsum board)
13.5

14.5

15.5

16.5

17.5

18.5

19.5

20.5

21.5

22.5

23.5

24.5

25.5

Effective R-value if insulation is installed in cavity between metal framing (includes gypsum board)
na
na
na
na
3.3
3.8
4.3
4.7
5.1
5.4
5.8
6.1

na
na
na
na
3.3
3.9
4.4
4.8
5.2
5.6
5.9
6.3

na
na
na
na
3.4
3.9
4.4
4.9
5.3
5.7
6.1
6.4

na
na
na
na
3.4
4
4.5
5
5.4
5.8
6.2
6.6

na
na
na
na
na
4
4.6
5.1
5.5
5.9
6.3
6.7

na
na
na
na
na
4.1
4.6
5.1
5.6
6
6.5
6.8

na
na
na
na
na
4.1
4.7
5.2
5.7
6.1
6.6
7

na
na
na
na
na
4.1
4.7
5.2
5.8
6.2
6.7
7.1

na
na
na
na
na
na
4.8
5.3
5.8
6.3
6.8
7.2

na
na
na
na
na
na
na
5.4
5.9
6.4
6.8
7.3

na
na
na
na
na
na
na
5.4
5.9
6.4
6.9
7.4

na
na
na
na
na
na
na
5.4
6
6.5
7
7.5

na
na
na
na
na
na
na
5.5
6
6.6
7.1
7.6

1 ASHRAE 90.1-2007, www.ASHRAE.org

wall criteria. These criteria generally allow a lower wall
R-value. The ANSI/ASHRAE/IESNA Standard 90.1-2007
requires a heat capacity greater than 7 Btu/ft2·°F, except lightweight concrete walls with a unit weight not

greater than 120 pcf need only have a heat capacity of
5 Btu/ft2·°F or greater. Table 5.3.9 provides heat capacities of concrete walls. These walls meet the minimum
requirements for mass wall criteria in almost all cases.

ARCHITECTURAL PRECAST CONCRETE

|

411

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.4 Heat Capacity / 5.3.5 Thermal Mass

Table 5.3.9 Heat Capacity of Concrete.

Heat Capacity,
Btu/ft2·°F

Concrete
Thickness,
in.

145 pcf

110 pcf

3

7.2

5.5

4

9.6

7.3

5

12.0

9.2

6

14.4

11.0

7

16.8

12.8

8

19.2

14.6

9

21.6

16.5

10

24.0

18.3

11

26.4

20.2

12

28.8

22.0

ASHRAE 90.1-2007, www.ASHRAE.org

5.3.5 T hermal Mass
The thermal mass provided by concrete buildings saves
energy in many climates. Thermal mass shifts peak
loads to a later time and reduces peak energy requirements for building operations. Laboratory, analytical,
and field studies support this concept. Thermal resistance (R-values) and thermal transmittance (U-factors),
discussed in Section 5.3.3, do not take into account
the effects of thermal mass, and by themselves, are
inadequate in describing the heat transfer properties
of construction assemblies with significant amounts of
thermal mass.
As previously discussed, common thermal properties
of construction materials and air spaces are based on
steady state tests, which measure the heat that passes
from the warm side to the cool side of the test specimen. Thermal transmittance (U-factor) and its reciprocal, overall R-value is generally considered the most
significant indication of heat gain because low mass
buildings constructed of metal or wood frame have
heat losses proportional to the overall area-weighted
U-factor of the building envelope (walls and roof). Also,
U-factors and R-values are relatively easy to calculate
since they are based on steady-state conditions.
However, the steady-state condition is rarely realized
in actual practice. External conditions (temperatures,
position of the sun, presence of shadows, etc.) vary
throughout a day, and heat gain is not instantaneous

412

|

ARCHITECTURAL PRECAST CONCRETE

through most solid materials, resulting in the phenomenon of time lag (thermal inertia). As temperatures rise
on one side of a wall, heat begins to flow toward the
cooler side. Before heat transfer can be achieved, the
wall must undergo a temperature increase. The thermal energy necessary to achieve this increase is related
to heat capacity.
Due to its density, concrete has the capacity to absorb
and store large quantities of heat. This thermal mass
allows concrete to react very slowly to changes in outside temperature. This characteristic of thermal mass
reduces peak heating and cooling loads and delays the
time at which these peak loads occur by several hours
(Fig. 5.3.17[a]). Mass effects vary with climate, building type, orientation, position of mass within the wall,
and other factors, so quantifying their effects is more
challenging than calculating R-values. Mass effect,
glass area, air infiltration, ventilation, building orientation, exterior color, shading or reflections from adjacent structures, surrounding surfaces or vegetation,
building shape, number of stories, wind direction, and
speed all affect energy use.
Analytical and experimental studies have shown
that the use of materials with thermal mass in buildings reduces heating and cooling peak loads, and
thus reduces equipment size compared with buildings
constructed with lightweight materials. Small HVAC
equipment that runs continuously uses less energy
than large equipment that is run intermittently as it responds to peak loads. By lowering peak loads, energy
is saved. Peak cooling loads in office buildings are often in midafternoon. Properly designed thermal mass
can shift a portion of the load and undesireable heat
gain from midafternoon until later when the building
is unoccupied or when peak load electricity costs are
less. Also thermal mass on the interior building surface
will help absorb heat gains in the office space.
Energy use differences between light and heavy materials are illustrated in the hour-by-hour computer
analyses shown in Fig. 5.3.17. Fig. 5.3.17(a) compares
the heat flow through three walls having the same Ufactor, but made of different materials. The concrete
wall consisted of a layer of insulation sandwiched between inner and outer wythes of 2 in. concrete with a
combined weight of 48.3 psf. The metal wall, weighing 3.3 psf, had insulation sandwiched between an exterior metal panel and 1/2 in. drywall. The wood frame
wall weighed 7.0 psf and had wood siding on the outside, insulation between 2 × 4 studs, and 1/2-in. drywall

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.5 Thermal Mass

Fig. 5.3.17(a-c) Heating and cooling load comparisons.

3

All Walls have
U-Values = 0.091

Heating Load,
BTU/Hr. FT2

Concrete Sandwich
Wall (2" Inner and
Outer Wythes)

2
1
0

2

4

6

Time
8 10 12 14 16 18 20 22 24 Hr.

1
Metal Wall
2
3

3

Wood Frame Wall
Cooling Load, BTU/Hr. FT2

All Walls have
U-Values = 0.091

Heating Load,
BTU/Hr. FT2
2"
4"

2

6"

1
0

(a) Walls

8"

2

4

6

8 10 12 14 16 18 20 22 24 Time
Hr.

1
2
Normal Weight Concrete
3

3

Cooling Load, BTU/Hr. FT2
Heating Load,
BTU/Hr. FT2

All Walls have
U-Values = 0.091
2" Concrete Inner Wythe

2

4"
6"
8"

1
0

(b) Concrete Walls

2

4

6

Time
8 10 12 14 16 18 20 22 24 Hr.

1
2
3

Cooling Load, BTU/Hr. FT2

(c) Concrete Sandwich Walls

5

on the inside. The walls were exposed to simulated
outside temperatures that represented a typical spring
day in a moderate climate. The massive concrete wall
had lower peak loads by about 13 percent for heating
and 30 percent for cooling than the frame or non-mass
walls. Actual results for buildings depend on the location, time of year, and building design.
Concrete walls of various thicknesses that were exposed to the same simulated outside temperatures are
compared in Fig. 5.3.17(b). The walls had a layer of
insulation sandwiched between concrete on the outside and 1/2 in. drywall on the inside; U-factors were
the same. The figure shows that the more massive the
wall, the lower the peak loads and the more the peaks
were delayed.
Fig. 5.3.17(c) compares concrete sandwich panels having an outer wythe of 2 in., various thicknesses of insulation, and various thicknesses of inner wythes. All walls
had U-factors of 0.091 and were exposed to the same
simulated outside temperatures. The figure shows that
by increasing the thickness of the inner concrete wythe,
peak loads were reduced and delayed.
ASHRAE Standard 90.1 acknowledges the thermal
mass benefits of concrete walls in specifying lower
minimum insulation R-value and higher maximum wall
U-factors for mass (concrete) wall construction. For
example, in a region with 5401-7200 heating degree
days base 65°F (HDD65 [Chicago]), the minimum Rvalue for concrete wall insulation is R 7.6ci (ci = continuous insulation) and for steel framed walls the minimum R-value for the wall insulation is R 13 + R 3.8 ci.
For the same region the maximum wall U-factor for
concrete walls is 0.123 and for steel framed walls the
maximum U-factor is 0.084.
In fact, research conducted by Oak Ridge National
Laboratory (ORNL) on the computer modeling and simulation of dynamic thermal performance of insulated
concrete walls versus traditional wood frame shows
that insulated concrete sandwich walls constructed with
composite connector technology utilizes the thermal
mass effect of concrete to create an “equivalent wall
performance R-value” several times greater than what is
estimated by a traditional material R-value calculation.6
In this study, six climates were evaluated – Atlanta,
Denver, Miami, Minneapolis, Phoenix, and Washington,
D.C. Of these cities, the difference was most dramatic
6 “Thermal Performance of Prefabricated Concrete Sandwich Wall Panels,” J.
Kosny, P. Childs and A. Desjarlais, Oak Ridge National Laboratory Buildings
Technology Center, October 2001

ARCHITECTURAL PRECAST CONCRETE

|

413

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.5 Thermal Mass

Energy saving benefits of thermal mass are most
pronounced when the outside temperature fluctuates above and below the balance temperature of the
building, causing a reversal of heat flow within the
wall. The balance point is generally between 50 and
70ºF, depending on the internal gains due to people,
equipment and solar effects. These ideal conditions for
thermal mass exist on a daily basis at all locations in the

in Phoenix, where a comparable R-value of conventional
wood frame exterior wall would need to be 2.9 times
higher than the steady state R-value of an insulated concrete sandwich panel wall to produce the same energy
loads. Therefore a comparative wood frame wall R-value
would need to have an R 31 to achieve the same effect
as an R 11 insulated concrete sandwich panel wall constructed with composite connector technology.

Table 5.3.10 (a) Design Considerations for Building with High Internal Heat Gains1.

1

2

Reduce
Infiltration

2

Daylighting

2

2

3

3

2

3

3

Dark

Climate Classification

Surface Color
Light

External
Fins3

1

Thermal
Mass

Increase
Insulation

Relative Importance of Design Considerations2

Winter
Long heating
season
(6000 degree
days or more)
Moderate heating
season
(3000–6000
degree days)
Short heating
season
(3000 degree
days or less)

With sun and wind

4,5

With sun without wind

1

3

3

Without sun with wind

Without sun and wind
1

2
2

2

1

3

3

With sun and wind

2

2

1

1

2

2

With sun without wind

2

2

1

2

2

Without sun and wind

1

2

2

2

Without sun with wind

1

2

2

2

With sun and wind

3

1

1

1

With sun without wind

3

1

1

1

Without sun and wind

2

1

1

1

Without sun with wind

2

1

1

1

1

Summer
Long cooling
season
(1500 hr. at 80 ºF)

Dry or humid

3

3

3

2

3

Moderate cooling
season (600–
1500 hr. at 80 °F)

Dry or humid

3

2

2

2

3

Short cooling
season
(Less than 600 hr.
80 °F)

Dry or humid

2

1

1

2

3

1 Includes office buildings, factories, and commercial buildings.
2 Higher numbers indicate greater importance.
3 Provide shading and protection from direct wind.
4 With sun: sunshine during at least 60% of daylight time.
5 With wind: average wind velocity over 9 mph.

414

|

ARCHITECTURAL PRECAST CONCRETE

IMPORTANCE
RATING KEY
3 High
2 Medium
1 Low

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.5 Thermal Mass

United States and Canada during at least some months
of the year. Thermal mass is most effective in conserving energy in the sun-belt regions in the Southern and
Western United States, because these daily temperature fluctuations occur throughout the year. Thermal
mass also works well when daily temperatures have
large variations between the daytime high and night-

5

time low and when outdoor air can be used for nighttime ventilation. These conditions are most prevalent
in the western states. Designs employing thermal mass
for energy conservation should be given a high priority
in these areas.
Another factor affecting the behavior of thermal mass

Table 5.3.10 (b) Design Considerations for Building with Low Internal Heat Gains1.

Reduce
Infiltration

2

Dark

3

Surface Color
Light

External
Fins3

Climate Classification

Increase
Insulation

Thermal
Mass

Relative Importance of Design Considerations2

Winter
Long heating
season
(6000 degree
days or more)
Moderate heating
season
(3000–6000
degree days)
Short heating
season
(3000 degree
days or less)

With sun and wind

4,5

3

3

With sun without wind

3

3

3

Without sun and wind

3

2

3

Without sun with wind

3

2

2

3

With sun and wind

1

2

1

2

3

With sun without wind

1

2

2

3

2

1

3

1

3

Without sun and wind
Without sun with wind

1

2

With sun and wind

2

1

1

2

With sun without wind

2

1

1

2

Without sun and wind

1

1

2

1

1

2

Without sun with wind

1

Summer
Long cooling
season
(1500 hr. at 80 ºF)
Moderate cooling
season (600–
1500 hr. at 80 °F)
Short cooling
season
(less than 600 hr.
at 80 °F)

Dry6 or humid7

3

2

2

3

Dry

2

1

1

2

Humid

2

1

1

3

Dry or humid

1

1

1 Includes low-rise residential buildings and some warehouses.
2 Higher numbers indicate greater importance.
3 Provide shading and protection from direct wind.
4 With sun: sunshine during at least 60% of daylight time.
5 With wind: average wind velocity over 9 mph.
6 Dry: daily average relative humidity less than 60% during summer.
7 Humid: daily average relative humidity greater than 60% during summer.

IMPORTANCE
RATING KEY
3 High
2 Medium
1 Low

ARCHITECTURAL PRECAST CONCRETE

|

415

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.5 Thermal Mass

is the availability of internal heat gains. This includes
heat generated inside the building by lights, equipment, appliances and people. It also includes heat from
the sun entering through windows. Generally, during
the heating season, benefits of thermal mass increase
with the availability of internal heat gains; Tables
5.3.10(a) and 5.3.10(b) may be used as a guide. Thus,
office buildings which have high internal heat gains
from lights, people, and large glass areas represent an
ideal application for thermal mass designs. This is especially true if the glass has been located to take maximum advantage of the sun. During the cooling season,
thermal mass “coupled” or exposed to the building
occupied spaces will absorb internal gains, thereby
shifting the peak cooling periods. Concrete exposed to
the interior and not covered by insulation and gypsum
wallboard is best able to absorb internal gains, thereby
saving cooling energy.

The first phase of a botanical center used the high
mass characteristics of precast concrete to store heat
and stabilize temperatures (Fig. 5.3.18). The walls
consist of 12-in. sandwich panels having a 3-in. outer
wythe, 3 in. of insulation, and a 6 in. inner wythe resulting in an R-value of 16. The inside 6 inch layer of
concrete provided approximately 480,000 pounds of
mass for storage of passive solar heat. The high mass
radiates heat back into the structure in the late afternoon and evening. Precast concrete was also used for
its light color and its ability to reflect sunlight into the
garden area.
Only computer programs such as DOE-27, Energy-108,
and EnergyPlus9 that take into account hourly heat
transfer on an annual basis (8760 hours) are adequate
7 DOE-2 http://simulationresearch.LBL.gov
8 Energy-10, www.sbicouncil.org
9 Energy Plus, www.energyplus.gov.

Fig. 5.3.18
Quad City Botanical Center, Rock Island, Illinois; Architect: Change-Environmental Architecture; Photo: Dale Photographics, Inc.

416

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.5 Thermal Mass / 5.3.6 Condensation Control

in determining energy loss in buildings with mass walls
and roofs.
Building codes and standards provide prescriptive and performance paths for meeting requirements using thermal mass. Prescriptive paths have
required minimum or maximum values in easy-touse tables for each building component. Generally,
R-value requirements for mass walls are less than those
for wood or steel frame walls. To obtain a range of Rvalues, the precast concrete walls may have insulation
applied to the interior or the insulation may be fully
incorporated into a sandwich wall panel.
Performance paths are used to trade one energy saving measure for another. For instance, if the wall insulation does not meet the prescriptive requirements,
but the ceiling insulation exceeds the prescriptive requirements, then using a performance method may
show compliance of the whole building with the code.
Prescriptive paths are commonly used for typical buildings in states with newly adopted codes. Once designers become familiar with performance software, these
become more popular. Some performance methods
can be used to show energy savings beyond code,
and are used for sustainability programs or state tax
credits.
The performance paths in energy codes generally allow the use of an easy-to-use computer trade-off program or a detailed energy budget method. Generally
the more complicated the compliance tool, the more
flexibility the designer is allowed. Trade-off tools also
allow for innovation in design and materials. ENVSTD
is an easy-to-use program for determining compliance
of the building envelope of commercial buildings with
ASHRAE 90.1. COMcheck-EZ™ (www.EnergyCodes.
gov) is an easy-to-use program for determining commercial building compliance for ASHRAE 90.1, IECC
(www.IccSafe.org) and many state codes.
COMcheck-Plus™ is a more detailed program using
the whole building approach to determine compliance.
This program is useful when buildings have special
features such as large skylight areas. A detailed computer-based energy analysis program such as DOE2 or
Energy Plus calculate yearly energy consumption for a
building on an hourly basis. Such programs are useful
when using the energy budget method because other simpler compliance tools do not take into account
special features of the building or its components. The

5

energy budget method compares the annual energy
use of a building that meets prescriptive requirements
with the proposed building to determine compliance.
Codes provide rules and guidelines for the energy budget method. All of these performance path methods
incorporate thermal mass effects.
Energy codes often specify insulation requirements for
mass walls based on whether the insulation is on the
interior of the wall, integral or on the exterior. Interior
insulation isolates the mass from the interior, reducing
the ability of the thermal mass to moderate the indoor
temperature. Integral insulation refers to thermal mass
on both sides of the insulation, as with an insulated
sandwich panel wall. It should be noted that regardless
of insulation placement, insulated mass walls combine
the benefits of insulation and mass and are often quite
energy efficient.

5.3.6 Condensation Control
Moisture which condenses on the interior of a building
is unsightly and can cause damage to the building and
its contents. Even more undesirable is the condensation
of moisture within a building wall where it is not readily
noticed until damage has occurred. Moisture accumulation can cause wood to rot and metal to corrode.
Fungi and biological growth such as molds have the
potential to grow in the presence of moisture or at
relative humidities on the wall surface of 70% or higher. In general a favorable combination of the following conditions are required for growths to germinate,
sprulate, and grow:
1. Fungal spores settling on the surface
2. Oxygen availability
3. Optimal temperatures (40 to 100°F)
4. Nutrient availability
5. Moisture (liquid or vapor above 70%RH)
Although concrete does not provide nutrients for mold
growth, nutrients may be abundant as dirt and dust particles on the surface of the concrete. The first four conditions are met in almost every building. So, the primary
method in controlling biological growth is to avoid high
humidities and surface condensation. The key is to manage moisture by adhering to sound construction practices that minimize the potential for condensation.

ARCHITECTURAL PRECAST CONCRETE

|

417

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6 Condensation Control / 5.3.6.2 Sources of moisture

Guidance in this chapter to eliminate condensation
and prevent mold is from three recognized sources.10, 11,
12
and can be summarized as follows
1. Increase surface temperature or reduce moisture
level in the air.
2. Install a vapor retarder or vapor resistant material
on the inside of insulation in cold climates.
3. Install a vapor resistant material on the outside of
insulation in warm climates.
4. Prevent or reduce air infiltration.
5. Prevent or reduce rainwater leakage.
6. Pressurize or depressurize the building, depending
on the climate, so as to prevent warm, moist air
from entering the building envelope.

The different climate types are defined on the map in
Fig. 5.3.19 and described in Table 5.3.11.

5.3.6.2 Sources of moisture
Moisture can enter building walls from the interior,
exterior, soil, or the building materials themselves.

5.3.6.1 C limates

Interior sources of moisture include people, kitchen
and restroom facilities, and industrial processes. The
average person produces 2.6 pints per day through respiration and perspiration. This amount increases with
physical activity. Nearly all of the water used for indoor
plants enters the indoor air. Five to seven small plants
release 1 pint per day of water. In residential facilities,
a shower can contribute 0.3 pints per minute and a
kitchen 5 pints per day for a family of four. Active vents
that remove moist indoor air to the outdoors should be
provided in showers and kitchens.

Buildings in drier climates generally have less condensation problems than those in more humid climates.
Generally the U.S. can be divided into humid and dry
by a north-south line drawn through the center of the
state of Texas. Areas east are humid and those west are
dry. The exception is the northwest, where the coast of
Washington and Oregon are also humid; these locations are called “marine.” In drier climates, moisture
that gets on or into walls will tend to dry to the inside
10 ASHRAE Handbook of Fundamentals - 2001, American Society of
Heating, Refrigerating, and Air-Conditioning Engineers, Atlanta, GA,
Chapters 23, 24, and 25. www.ASHRAE.org
11 Treschsel, Heinz, Moisture Control in Buildings, Publication No. MNL 18,
ASTM, West Conshohocken, PA, 1994. www.ASTM.org
12 Treschsel, Heinz, Moisture Analysis and Condensation Control in Building
Envelopes, Publication No. MNL 40, ASTM, West Conshohocken, PA,
2001. www.ASTM.org

|

However, even though buildings are more forgiving
in drier climates, condensation has the potential to occur in warm, cold, or mixed climates if walls are not
properly designed.

Good quality concrete is not damaged by moistureconcrete walls actually gain strength if they stay moist.

Causes of condensation are predominantly climate
dependent. The first cause occurs when outside conditions are cold and is due to moist interior air condensing on cold surfaces; locations with these conditions
will be called “cold.” The second cause occurs when
outside conditions are warm and humid and is due to
humid air entering the building and condensing on
cooler surfaces; locations with these conditions will be
called “warm.” Generally either of these conditions requires weeks rather than a few days for problems to
occur. Some locations experience long enough warm
and cold seasons to develop both types of condensation; these climates will be called “mixed.”

418

and outside more readily than in more humid climates.
For instance, when The Disney Company built Disney
World in Orlando in the 1970s, many of the structures
were constructed of the same painted wood construction and practices prevalent in Disneyland in southern
California. These structures did not hold up well in the
warm humid climate of central Florida.

ARCHITECTURAL PRECAST CONCRETE

Industrial processes, storage of moist materials,
swimming pools, commercial laundries, kitchens, and
ice rinks all contribute to indoor sources of moisture.
Buildings with these conditions should be designed for
the particular moisture conditions anticipated. In all
cases, guidelines of ANSI/ASHRAE Standard 6213 should
be followed for proper ventilation of indoor air.
Outdoor sources include precipitation and infiltration. Rain and melting snow cause problems when
the ground against walls is not pitched to move water away, or when plants that require frequent watering are located near walls. Vegetation near buildings
should be able to survive without watering or a buffer
area of decorative gravel can be placed. Landscaping
13 ANSI/ASHRAE Standard 62-2001 – Ventilation for Acceptable Indoor Air
Quality, American Society of Heating, Ventilating, and Air-Conditioning
Engineers (ASHRAE), Atlanta. http://www.ashrae.org

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.2 Sources of moisture

5

Fig. 5.3.19 Climate zones for moisture.

ASHRAE 90.1-2007

Table 5.3.11 Climate Zones for Moisture.

Zone No.

Description

Representative U.S. Cities

1A, 2A, and 3A south
of the humid line

Warm, humid

Miami, FL; Houston, TX

2B

Warm, dry

Phoenix, AZ

3A north of the
humid line, 4A

Mixed, humid

Memphis, TN; Baltimore, MD

3B, 3C, 4B

Mixed, dry

El Paso, TX; San Francisco, CA;
Albuquerque, NM

4C*

Cool, marine

Salem, OR

*

Cold, humid

Chicago, IL; Burlington, VT;

*

5A, 6A
5B, 6B

Cold, dry

Boise, ID; Helena, MT

*

7

Very cold

Duluth, MN

8

*

Subarctic

Fairbanks, AK

*For Canadian locations, climate zones are defined on the basis of Heating Degree Days Base 65 °F (HDD65F):
Zone 4C: 3600 < HDD65F ≤ 5400
Zone 5: 5400 < HDD65F ≤ 7200
Zone 6: 7200 < HDD65F ≤ 9000
Zone 7: 9000 < HDD65F ≤ 12,600
Zone 8: 12,600 < HDD65F

ARCHITECTURAL PRECAST CONCRETE

|

419

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.2 Sources of moisture / 5.3.6.3 Condensation on surfaces

near buildings has led to automatic sprinkler systems
that “water” building walls. Moisture from precipitation should be controlled to prevent it from entering
the walls or building. A primary and secondary line of
defense should always be provided. For instance if joint
sealant is used to prevent precipitation from entering a
wall, a second line of joint sealant should be provided
behind the first to keep out moisture should the first
deteriorate.
Infiltration of moist air is caused by several sources.
Due to the stack effect in buildings (warm air rises),
outdoor air enters the building through cracks and
joints near the bottom of the building and exits near
the top. This effect is greater for taller buildings. Also,
heating and cooling systems should have adequate air
intake systems. Otherwise when the system is operating and exhausting air, it will depressurize the building
and air can be drawn into the building through cracks,
joints, and building materials. When the moisture content of outdoor air is greater than the indoor air, for
example in warm humid climates, infiltration and depressurization bring moisture into the building. Moist
air also enters the building through cracks, joints, and
building materials when the vapor pressure of the
outdoor air is greater than the indoor air. Again, this
occurs on warm humid days or cooler days with high
relative humidity.
Soil has the potential to provide a continuous supply
of moisture to concrete through slabs and foundations.
Capillary breaks between the foundation and above
grade walls can reduce this potential. The ground
should be sloped away from buildings and adequate
drainage and waterproofing should be provided. As
land becomes more scarce and costly, more buildings
are being built on less desirable sites that previously
ponded water; drainage must be properly considered
in these areas. Also, any water draining from adjacent
sites onto the subject building site needs to be properly channeled away from buildings. Vapor retarders
should be installed beneath all concrete floor slabs in
direct contact with the concrete to prevent moisture
from moving up into the building. The vapor retarder
should be installed above a granular subbase layer and
directly beneath the concrete slab.
Building materials contribute significantly to moisture inside buildings, known as “moisture of construction,” during the first years after construction.

420

|

ARCHITECTURAL PRECAST CONCRETE

Concrete contributes significant moisture since it starts
as a saturated material. Precast concrete dries during
storage and continues to dry in the built structure until
the pores near the surface reach an equilibrium moisture content with the indoor air. Wood and materials
stored outdoors are also contributors. Many buildings have noticeable condensation the first year after
construction that will subside in subsequent years.
Dehumidification and adequate ventilation can help alleviate condensation due to moisture of construction.

5.3.6.3 Condensation on surfaces
Causes. Condensation occurs on surfaces inside
buildings when the surface temperature is less than the
dew point of the indoor air. The dew point of the air
depends on its relative humidity. Dew-point temperatures to the nearest ºF for various temperatures and
relative humidities are shown in Table 5.3.12. In the
summer in humid climates the relative humidity (RH) of
the indoor air is generally in the range of 50 to 80%. In
the winter in cold climates the relative humidity of the
indoor air is generally in the range of 20 to 40%.
Relative humidity is the ratio of the amount moisture
in the air to the amount of moisture the air can hold
(saturation). Colder air holds less moisture. In climates
like Chicago the average relative humidity outdoors
averages approximately 70%. Yet, the amount of
moisture in the outdoor air is much less in the winter
because the air holds less moisture. When this drier air
is brought inside and heated up, the resulting relative
humidity at 70°F is low; often in the range of 15 to
25%.
Example 5.3.3 – Condensation on a beverage
can. Condensation may occur on a beverage can inside of a 75ºF building during the summer, but not at
the same temperature in the winter. In the summer at
75ºF and 80%RH, the dew point is 68ºF. If the temperature of the can is less than 68ºF, condensation will
occur on the can. In winter at 75ºF and 30%RH, the
dew point is 42ºF. If the can is less than 42ºF condensation will occur on the can. Also, at the low RH in the
winter, moisture that would condense on the can will
evaporate quickly and may not be noticed.
Condensation on surfaces occurs most frequently due
to cool indoor surface temperatures or high indoor humidity levels. These can be the result of many factors:
1. Inadequate heating and ventilation can result in

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.3 Condensation on surfaces

Δtn = Rn · (ti – to)/ RT

Table 5.3.12 Dew-Point Temperatures1.

Dry Bulb
or Room
Temperature, °F

Equation 5.3.5

where:

Relative Humidity (RH), %

Δtn = temperature gradient or drop
through material “n”

10

20

30

40

50

60

70

80

90

100

40

-8

5

13

19

24

28

31

34

37

40

45

-4

9

17

23

28

32

36

39

42

45

50

-1

13

21

27

32

37

41

44

47

50

55

3

17

25

31

37

41

45

49

52

55

ti

= indoor air temperature

60

6

20

29

36

41

46

50

54

57

60

to

= outdoor air temperature

65

10

24

33

40

46

51

55

59

62

65

70

13

28

37

45

51

55

60

64

67

70

RT = thermal resistance of wall including air film resistances

75

17

31

42

49

55

60

65

68

72

75

80

20

36

46

54

60

65

69

73

77

80

85

24

39

50

58

64

70

74

90

27

44

54

62

69

74

79

Rn = thermal resistance of material
“n”

2. Furniture or partitions placed up against walls may
prevent adequate heating or air flow and produce
cool surfaces.

The calculation of the temperature
gradient profile through a wall assem78 82 85 bly due to a temperature difference
83 87 90 between indoors and outdoors can be
used to determine whether there may
be a problem with condensation or differential thermal movement. The temperature gradient
alone is not sufficient to accurately locate the dew point
within the assembly but it can be used as a guide for
determining where condensation may occur from exfiltrating or infiltrating air. The assumption of steady-state
conditions in this method is seldom satisfied due to fluctuations in temperatures within the wall. Nevertheless,
the calculation is useful to flag potential problems.

3. Closets, which are rarely conditioned, can also have
inadequate ventilation and cool surfaces.

Examples are provided for condensation on a cool surface in winter and summer.

4. Insufficient, damaged, or wet wall insulation can
cause cool surfaces.

Example 5.3.4 – Winter surface condensation due
to inadequate heat or air distribution
Assume that, due to poor air circulation, the indoor
air conditions are 75ºF and 30%RH near the top of the
wall and 40ºF with an equal amount of moisture in the
air near the bottom. This example is the same as the
beverage can, Example 5.3.3; condensation will occur
if the temperature of the wall is less than 42ºF. This
can be prevented by providing adequate heating and
ventilation along the full height of all walls.

1 Temperatures are based on a barometric pressure of 29.92 in. Hg.

cooler surface temperatures near the bottom of
walls. Heating must be provided near floor level or
with enough circulation to heat the lower portion
of rooms.

5. Thermal bridges, or areas of the wall that are not
insulated as well as others, can also produce cooler
surface temperatures.
6. High humidity caused by swimming pools, ice rinks,
or industrial processes can cause condensation on
indoor surfaces.
7. Cold air from air-conditioners blowing in the region of warm humid air can cause condensation
on indoor surfaces.
The potential for condensation can be determined
if wall temperatures and relative humidity of the air
are known. The temperature gradient through any
portion of a wall is directly proportional to its thermal
resistance. Therefore, the temperature gradient Δtn
through a material with a thermal resistance Rn can be
calculated using Equation 5.3.5:

5

Example 5.3.5 – Winter surface condensation due
to not enough insulation
Assume the indoor air conditions are 70ºF and
35%RH and the average outdoor temperature for the
day is 20ºF. Assume the wall is an insulated concrete
sandwich panel from the previous thermal resistance
calculation, Example 5.3.1. Compare this to a wall with
no insulation. First we will determine the temperatures
of the wall with insulation.

ARCHITECTURAL PRECAST CONCRETE

|

421

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.3 Condensation on surfaces

side the room is 65ºF while that of the uninsulated
wall surface is 39ºF. The surface film
resistance plays a much larger role in
Temp.
R Winter
Temp., ºF
an uninsulated wall. The temperature
Difference, ºF
gradient across the inside air film is 5ºF
Surface, outside air film
0.17
1
20
for the insulated wall and 31ºF for the
uninsulated wall. The dew point of air
Concrete, 2 in. (145 pcf)
0.13
1
21
at 70ºF and 35%RH is 42ºF. Since the
EPS insulation
inside surface of the uninsulated wall
6.00
42
22
(1.25 pcf), 11/2 in.
is 39ºF, condensation will form on the
inside surface.
Concrete, 2 in. (145 pcf)
0.13
1
64

Thermal Resistance and Temperatures of Insulated Wall.

A.
B.
C.
D.
E.

Surface, inside air film

0.68

5

Total

7.11

50

U = 1/R

0.14

The thermal resistance of the wall, RT, equals 7.11.
The temperature difference across the wall, ti – to,
equals 70ºF – 20ºF = 50ºF. The temperature difference
across any layer is calculated using Equation 5.3.5.
The temperature difference across the air film equals
0.17(50) / 7.11 or 1ºF. The remaining temperature differences are calculated in the same manner as shown
on the previous page. The temperature differences
Thermal Resistance and Temperatures of Uninsulated Wall.

R Winter

• Junctions of floors and walls, walls
and ceilings, walls and roofs

A.

Surface, outside air film

0.17

8

20

• Around wall or roof openings

B.

Concrete, 4 in. (145 pcf)

0.25

11

28

• At perimeters of slabs on grade

C.

Surface, inside air film

0.68

31

39

• At connections, if insulation is
penetrated

Total

1.10

50

70

U = 1/R

0.91

Note that the air temperature of the room is 70ºF
and the temperature of the insulated wall surface in-

|

Thermal bridges, such as a full thickness of concrete
along panel edges, will behave similar to the uninsulated wall in Example 5.3.5. Thermal bridges may also
occur at;

Temp.
Temp., ºF
Difference, ºF

are subtracted from the indoor air temperature (or
added to the outdoor temperature) to determine temperatures at boundaries between materials and are
shown above in the right column. The inside surface
of the wall, between the concrete and the inside air
film, is 65ºF.
Shown above is the determination of the thermal resistance and temperatures of an uninsulated wall.

422

Also note that the average outdoor
air temperature for the day was used
70
in calculations. This average rather
than the lowest daily temperature was
used for two reasons. First, thermal
mass of the concrete will tend to moderate the indoor surface temperature
so that using an extreme temperature expected for just
a few hours may be too conservative. Secondly, if a
condensation occurrence is predicted for only a few
hours, it will often occur and evaporate without causing problems.
65

ARCHITECTURAL PRECAST CONCRETE

• Any place metal, concrete, or a
highly conductive material penetrates an insulation layer, such as
metal shear connectors

Condensation can develop at these locations especially if they are in corners or portions of a building that
receive poor ventilation.
Example 5.3.6 – Summer surface condensation
Condensation on wall surfaces also occurs in summer conditions. Cold air from air-conditioners blowing
in the region of warm humid air can cause condensation on indoor surfaces. This most frequently happens
when wall air-conditioner units are placed near win-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.3 Condensation on surfaces / 5.3.6.4 Condensation within walls and use of vapor retarders

dow or door frames that allow humid air to enter the
conditioned space.
Assume the average daily outdoor conditions are 80ºF
and 75%RH. Assume this air can enter a room in a
gap between the top of an air-conditioning unit and
the bottom of a window. Assume the air conditioning unit blows enough cool air in the vicinity of a wall
so that the wall surface temperature is 65ºF. Since the
dew point of the moist air is 71ºF, condensate will form
on the cool wall surface. This illustrates the need to
provide adequate joint sealing to prevent the entry of
humid air.
Prevention of Condensation on Wall Surfaces.
All air in buildings contains water vapor. If the inside
surface temperature of a wall is too cold, the air contacting this surface will be cooled below its dew-point
temperature and water will condense on that surface.
Condensation on interior room surfaces can be controlled both by suitable construction and by precautions such as: (1) reducing the interior RH or dew point
temperature by dehumidification equipment or ventilation; or (2) raising the temperatures of interior surfaces that are below the dew point, generally by use
of insulation.
The interior air-dew point temperature can be lowered by removing moisture from the air, either through
ventilation or dehumidification. Adequate surface temperatures can be maintained during the winter by incorporating sufficient thermal insulation, using double
glazing, circulating warm air over the surfaces, or directly heating the surfaces, and by paying proper attention
during design to the prevention of thermal bridging.

5.3.6.4 C ondensation within walls and
use of vapor retarders
Although condensation due to air movement is usually much greater than that due to vapor diffusion for
most buildings, the contribution from water vapor diffusion can still be significant. In a well-designed building, the effects of air movement and water vapor diffusion in walls and roofs are considered.
Vapor retarders. Air barriers (also called air retarders) and vapor retarders (also called vapor barriers) are
often confused. An air barrier is used to reduce the
amount of infiltration (or leakage) or exfiltration of air
into a conditioned space. A vapor retarder is used to
prevent, or more correctly greatly reduce, water vapor

5

(moisture) from moving through building materials.
A vapor retarder can be used as an air barrier. An air
barrier on the outside of a building in a cold climate
generally needs to let moisture escape, so should not
function as a vapor retarder. If the air barrier will also
be serving as a vapor retarder, or if it has a low permeance to vapor diffusion, then its position within the
building envelope must be carefully considered in relation to the other envelope components.
The principal function of a vapor retarder is to impede
the passage of moisture as it diffuses through the assembly of materials in a building envelope, to control
the location of the dew point in the assembly and to
ensure there is a manageable flow of moisture across
the assembly. The basic principles, simply stated are:
• Moisture migrates through building materials due
to a difference in temperature or RH or both between the inside and outside.
• Sometimes this moisture migration will cause condensation. The correct type and placement of insulation and a vapor retarder will prevent condensation on cold portions within a wall.
• The vapor retarder or vapor retarding materials are
generally placed on the side of the wall that is warm
most of the year.
• If a vapor retarder with low permeance is selected, the
materials on the opposite side should have higher
permeance so the wall is able to dry to that side, if
necessary.
These principles are covered in depth in the sections
that follow.
Most codes and references consider a material or
membrane with a permeance of 1 perm or less a vapor
retarder; less than 0.1 perms is considered vapor impermeable and between 0.1 and 1 perm is considered
semi-impermeable. Materials or membranes with a
permeance greater than 10 are considered permeable.
In the range of 1 to 10 perms, materials are considered
semi-permeable.
Concrete as a vapor retarder. Normalweight, quality concrete can be considered a semi-impermeable vapor retarder in thicknesses of 3 in. or more. Published
values of concrete permeability are approximately 3
perm·in., so that 3 in. of concrete has a permeance
of approximately 1 perm, provided it remains relatively
crack-free. Permeance is a function of the water-ce-

ARCHITECTURAL PRECAST CONCRETE

|

423

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Table 5.3.13 Typical Permeance (M) and Permeability (µ) Values1.

Material

M
µ
Perms Perm-in.

Concrete (1:2:4 mixture)2

—

3.2

Wood (sugar pine)

—

0.4 – 5.4

Extruded polystyrene (XPS)

1.2

Expanded polystyrene, bead (EPS)

—

2.0 – 5.8

Polyisocyanurate

—

4.0 – 6.6

Polyicynene

50

Glass fiber batt

120

Kraft paper

1

Plaster on gypsum lath (with studs)

20

Gypsum wallboard, 0.375 in.

50

Polyethylene, 2 mil

0.16

Polyethylene, 4 mil

0.08

Polyethylene, 6 mil

0.06

Aluminum foil, 0.35 mil

0.05

Aluminum foil, 1 mil

0.00

Built-up roofing (hot mopped)

0.00

Duplex sheet, asphalt laminated,
aluminum foil one side

0.0023

Paint
1 coat primer plus 2 coats latex
on gypsum wallboard

3 to 20

1 coat primer plus 2 coats acrylic
on gypsum wallboard

7

1 coat primer plus 2 coats
synthetic on gypsum wallboard

3

1 coat primer plus 2 coats oil on
gypsum wallboard

5

2 coats asphalt paint on plywood

0.4 3

2 coats enamel on smooth plaster

0.5
– 1.5

	Various primers plus 1 coat flat
oil paint on plaster

1.6
– 3.0

Breather type membrane

3 – 25

1 ASHRAE Handbook of Fundamentals and other sources. Values vary depending on the moisture content of the material.
2 Permeances for concrete vary on the concrete’s water-cement ratio and
other factors.
3 Dry-cup (ASTM E 96).

ment ratio of the concrete. A low water-cement ratio,
such as that used in most precast concrete members,
results in concrete with low permeance.
Where climatic conditions demand, insulation, sufficient concrete, or the addition of a vapor retarder is
generally necessary in order to prevent condensation.
Thicknesses of 1 in. or more of rigid extruded polystyrene board (XPS) or 2 to 3 in. of expanded polystyrene
(EPS), if properly applied, will serve as its own vapor
retarder. In such cases, for cold climates, the insulation
can be installed on a complete bed of adhesive applied
to the interior of the inner wythe of the wall with joints
fully sealed with adhesive, to provide a complete barrier to both air and vapor movement.
Codes. The International Energy Conservation Code
(IECC)14 requires a vapor retarder with of 1 perm or less
on the inside of insulation in cold climates. However it
allows for an exception where moisture or its freezing
will not damage the materials, or where other means
are provided to prevent condensation. This requirement is workable for concrete since 3 in. of concrete
has a perm of approximately 1 perm. The important
concepts are whether condensation will occur and, if it
does, will it damage the materials.
At present, the Massachusetts energy code is more
restrictive that the IECC.15 This code requires a vapor retarder of 0.1 perms on the indoor side of the insulation.
Concrete wall systems can generally meet the code under the exceptions that require calculations because the
condensing surface is the warm side of the insulation,
and the temperature at that surface is kept above the
dew point of the indoor air. This code also requires that
the materials and finishes on the outdoor side of the
insulation have permeances at least 10 times greater
than that on the inside. This requirement is needed to
allow the wall to dry to the outdoor side since the low
permeance will not allow it to dry to the indoor side.
Codes that have blanket requirements such as these
for all wall systems may cause more moisture problems
since low permeance materials sometimes prevent walls
from drying.
Other materials. Building materials have water vapor permeances from very low to very high, see Table
5.3.13. Actual values for a given material vary depending on the moisture content of the material. Two
14 International Energy Conservation Code (IECC), International Code
Council, Inc., Country Club Hills, IL, www.ICCsafe.org
15 Massachusetts Energy Code. www.mass.gov/bbrs/780 CMR Chapter 13.pdf.

424

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

commonly used test methods are the water method
(wet cup) and desiccant method (dry cup) methods in
ASTM E96, “Standard Test Methods for Water Vapor
Transmission of Materials.” Specimens are sealed over
the tops of cups containing either water or desiccant,
placed in a controlled atmosphere usually at 50%
relative humidity, and weight changes measured. The
change in weight represents the rate of moisture passing through the specimen.
When properly used, low permeance materials keep
moisture from entering a wall assembly. Materials with
higher permeance allow construction moisture and
moisture which enters inadvertently, or by design, to
escape.
When a material such as plaster or gypsum board
has a permeance which is too high for the intended
use, a vapor retarder can be used directly behind such

5

products. Polyethylene sheet, aluminum foil and building paper with various coatings are commonly used.
Proprietary vapor retarders, usually combinations of
foil and polyethylene or asphalt, are frequently used
in freezer and cold storage construction. When vapor
retarders are added sheets or coatings, they should be
clearly identified by the designer and be clearly identifiable by the general contractor.
Water vapor diffusion occurs when water vapor molecules diffuse through solid interior materials. The passage of water vapor through material is in itself generally not harmful. It becomes of consequence when,
at some point along the vapor flow path, a temperature level is encountered that is below the dew-point
temperature and condensate accumulates. The rate
of vapor movement is dependent on the permeability
of the materials, the vapor pressure, and temperature

Table 5.3.14 Water Vapor Pressures at Saturation (SVP) for Various Temperatures.

Temp.,
°F

SVP,
in. Hg

Temp.,
°F

SVP,
in. Hg

Temp.,
°F

SVP,
in. Hg

Temp.,
°F

SVP,
In. Hg

-30

0.007

17

0.089

38

0.229

59

0.504

-20

0.013

18

0.093

39

0.238

60

0.522

-10

0.022

19

0.098

40

0.248

61

0.541

-5

0.029

20

0.103

41

0.258

62

0.560

0

0.038

21

0.108

42

0.268

63

0.580

1

0.040

22

0.113

43

0.278

64

0.601

2

0.042

23

0.118

44

0.289

65

0.622

3

0.044

24

0.124

45

0.300

66

0.644

4

0.046

25

0.130

46

0.312

67

0.667

5

0.049

26

0.136

47

0.324

68

0.691

6

0.051

27

0.143

48

0.336

69

0.715

7

0.054

28

0.150

49

0.349

70

0.739

8

0.057

29

0.157

50

0.363

71

0.765

9

0.060

30

0.164

51

0.376

72

0.791

10

0.063

31

0.172

52

0.391

73

0.819

11

0.066

32

0.180

53

0.405

74

0.847

12

0.069

33

0.188

54

0.420

75

0.875

13

0.073

34

0.195

55

0.436

76

0.905

14

0.077

35

0.203

56

0.452

77

0.935

15

0.081

36

0.212

57

0.469

78

0.967

16

0.085

37

0.220

58

0.486

79

0.999

80

1.032

Note: 1 in. Hg = 0.491 psi. Actual vapor pressure = SVP x (%RH).

ARCHITECTURAL PRECAST CONCRETE

|

425

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

differentials. Generally, the greater the temperature
difference between inside and outside and the more
permeable the materials, the more vapor will travel
through the wall. Vapor pressures increase with temperature even if the relative humidities stay the same.
So, generally, the colder the outside temperature, the
greater the pressure of the water vapor in the warm
inside air compared to the cooler outside air. Water vapor pressures at saturation (100%RH) are provided in
Table 5.3.14. Leakage of moist air through small cracks
may be a greater problem than vapor diffusion.
Application. The location of the vapor retarder is
dependent on the wall construction and climate. A
solid precast concrete wall with appropriate joint sealant will act as a semi-impermeable vapor retarder in
many climates. If a separate air barrier membrane is
used, it should be clearly identified in the construction
documents, preferably on the drawings. While a vapor
retarder does not need to be perfectly continuous, care
should be taken to minimize the occurrence of small
discontinuities or imperfections such as unsealed laps,
cuts, and pin holes. The vapor retarder in a wall system
should be continuous from the floor to the underside
of the ceiling slab to prevent moisture from bypassing the vapor retarder. Wall penetration such as outlets
and window frames, should also be sealed.
Low-permeance paints, vinyl wall paper, or other similar materials that act as vapor retarders should not be
placed on the interior surface of concrete walls. Since
concrete acts also as a vapor retarder, an additional
vapor retarder prevents moisture within the wall from
evaporating.
Three common precast concrete systems and their
applicability for use in various climate zones (see Fig.
5.3.19) are presented in Fig. 5.3.20. These walls allow
concrete to dry without accumulating moisture within
the wall. The traditional practice for frame walls of
placing a vapor retarder behind gypsum wallboard in
cold climates is not recommended for these walls. The
recommendations were developed using typical indoor
relative humidities during winter for all building types.
Indoor relative humidities greater than these during
December, January, and February have the potential
to cause condensation within these or any wall/HVAC
system not properly designed.
The three walls in Fig. 5.3.20 are insulated to meet
the requirements of the 2004 International Energy

426

|

ARCHITECTURAL PRECAST CONCRETE

Conservation Code (IECC).16 The total wall including
the concrete, insulation, and interior finishes are considered in the design of a wall with low potential for
moisture problems. Providing insulation as required
by codes such as ASHRAE 90.1 or the IECC generally
provides cost effective levels of insulation for precast
concrete walls. Insulation requirements are dependent
on climate. The map in Fig. 5.3.19 is used to determine the climate zone number and letter required for
determining compliance with the IECC. The amount
of insulation required for the three walls is shown in
Fig. 5.3.20. For international locations, Appendix B of
ASHRAE 90.1-2004 provides tables with climate zone
numbers and letters. This appendix also provides the
climate zones in tabular form by U.S. county.
A precast concrete sandwich panel wall with concrete
on both sides of rigid insulation, Fig. 5.3.20(a), is recommended for Climate Zones 1 through 7 (all except
subarctic climates). Expanded polystyrene (EPS) or extruded expanded polystyrene insulation (XPS) may be
used. The insulation board shown in the wall details is
placed within the concrete during the precasting process prior to building construction. The overall thermal
resistance of a sandwich panel is greater (more energy
saving) if the ties connecting the concrete wythes are
plastic, composite fiberglass or epoxy coated carbon
grid rather than metal.
A precast concrete wall with continuous rigid insulation, Fig. 5.3.20(b), is recommended for Climate Zones
1 through 7 (all except subarctic climates). XPS insulation may be used in Climate Zones 1 through 7 and
EPS insulation may be used in Climate Zones 1 through
5. The lower permeance of the XPS is recommended
for the colder climates, Zones 6 and 7. The insulation
board should be applied continuously and in direct
contact with the precast concrete. This can be done
using adhesive, stick pins, or mechanical fasteners.
Continuous insulation uninterrupted by metal framing
is beneficial because metal framing reduces the effectiveness of fiberglass batt insulation and other insulation by more than half. For example, R13 insulation has
an effective R-value of 6 when placed between steel
frame members spaced 16 in. on center. The continuous insulation also reduces the potential for cold spots
on the interior and exterior surfaces caused by metal
framing. These can sometimes lead to condensation
16 International Energy Conservation Code (IECC), International Code
Council, Inc., Country Club Hills, IL, www.ICCsafe.org

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

5

‘Z’ furring or metal framing to create 1-in. minimum air
space (not required if no gypsum wallboard)
Precast Concrete Panel
/ -in. thick gypsum wallboard, painted (optional)

5 8

“A” Precast Sandwich Panel Wall
Not to scale

Extruded or Expanded Polystyrene Insulation, Continous
IECC Zone
Min. R-Value of Added Insulation
1, 2
None
3B
None
3A (Below “Warm Humid Line”)
None
3A (Above “Warm Humid Line”)
R7.5
3C, 4A, 4B
R7.5
4C, 5, 6
R10
7
R12.5

‘Z’ furring or metal framing to create 1-in. minimum air
space (optional)
Precast Concrete Panel

“B” Precast Concrete with Rigid Insulation
Not to scale

5 8
/ -in. thick gypsum wallboard, painted
(optional with no insulation)

Extruded or Expanded Polystyrene Insulation*,
Continous and Taped at Joints
IECC Zone
Min. R-Value of Added Insulation
1, 2
None
3B
None
3A (Below “Warm Humid Line”)
None
3A (Above “Warm Humid Line”)
R7.5
3C, 4A, 4B
R7.5
4C, 5, 6
R10
7
R12.5
*Expanded polystyrene insulation acceptable in climate zones 1-5

1-in. minimum air space
Precast Concrete Panel

5 8
/ -in. thick gypsum wallboard, painted
(optional with no insulation)

Batt Insulation with Kraft Paper on Inside Face,
in Zones 3, 4 and 5, in Metal Frame
IECC Zone
Min. R-Value of Added Insulation
1*, 2*
None
3B*
None
3A* (Below “Warm Humid Line”) None
3A (Above “Warm Humid Line”)
R13
3C*, 4, 5
R13
“C” Precast Concrete with Batt Insulation
Not to scale

*If insulation is used, kraft paper not recommended.
Only applicable for IECC zones 1-5

Fig. 5.3.20 Typical wall details.

ARCHITECTURAL PRECAST CONCRETE

|

427

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

and shadowing or other unsightly moisture problems
on the inside and outside surfaces of buildings. The
potential for shadowing in a sandwich panel wall is
less if the ties connecting the concrete wythes are plastic, fiberglass composite, or epoxy coated carbon grid
rather than metal.
Wood and steel frame walls have cavities where moisture can accumulate, causing wood to rot and metal to
corrode. The sandwich panel wall and concrete wall
with rigid insulation have no wall cavities within the
structural portion of the wall, thus reducing the possibility of unnoticed moisture accumulation and related damage. The only cavity is the air space between
the insulation and gypsum wallboard, if wallboard is
desired. This cavity is designed to keep the wallboard
dry. XPS insulation is particularly moisture resistant and
has low water absorption compared to other insulation materials while EPS has lower moisture absorption
compared to non-foam insulation materials.
A precast concrete wall with batt insulation (and
kraft paper where appropriate), Fig. 5.3.20(c), is recommended for Climate Zones 1 through 5. To prevent
potential moisture accumulation within the wall and
related problems, this type of wall construction is not
recommended for the colder climates, Zones 6, 7, and
8. The fiberglass insulation is installed between metal
framing. A 1-in. minimum air space is required between
the batts and the concrete to prevent the potential for
moisture to accumulate in the batt insulation. The air
space between the metal framing and the precast concrete reduces the potential for cold spots on the interior and exterior surfaces caused by the framing. These
can sometimes lead to condensation and shadowing
or other unsightly moisture problems on the inside
and outside surfaces of buildings. In Climate Zones
3A (above the warm humid line), 4, and 5, kraft-faced
batts are required to prevent condensation within the
walls during the winter.
The three walls in Fig. 5.3.20, with appropriate joint
sealant, will act as semi-impermeable vapor retarders
and allow concrete to dry without moisture accumulating within the walls. These constructions allow the
outside layer of concrete to dry to the outside and the
rest of the wall to dry to the inside. Latex paint with a
permeance of 5 to 10 perms on the drywall is generally
adequate. The sandwich panel wall and wall with rigid
insulation are assumed to have 11/2 to 2 in. of insulation in Zone 4, 2 in. in Zone 5, 2 to 21/2 in. in Zone 6,

428

|

ARCHITECTURAL PRECAST CONCRETE

and 21/2 to 3 in. in Zone 7. The wall with batt insulation
is assumed to have R13 fiberglass batts.
The location of the cold surfaces within a wall depends on the climate. Moisture generally moves into
wall systems from indoors when it is cold outside, and
into wall systems from outdoors when it is warm outside. Actual water vapor and moisture-laden air movement depends on the temperature and relative humidity indoors and outdoors, the moisture content of the
materials, and their absorption properties.
Cold Climates (Zones 5, 6, and 7). In these climates
the vapor retarding surface should be applied on or
near the warm side (inner surface) of assemblies. For
the concrete sandwich panel wall, the insulation, inside concrete wythe and painted gypsum wallboard,
if used, act as the semi-impermeable vapor retarder
during the winter. For the precast concrete wall with
rigid insulation, the insulation and painted gypsum
wallboard on the inside act as a semi-permeable vapor retarder during the winter. For the precast concrete
wall with batt insulation, the kraft paper and painted
gypsum wallboard act as a semi-permeable vapor retarder during the winter. For all three walls, the exterior concrete wythe acts as a semi-impermeable vapor
retarder during the summer. Providing an additional
low permeance vapor retarder on the inside of the wall
would create a “double vapor retarder” and prevent
moisture that accumulates within the wall from leakage or condensation from drying to the inside. For this
reason, a low permeance vapor retarder on the inside
of this wall system is not recommended.
For the sandwich panel wall and the precast concrete
wall with rigid insulation, the relative humidity of the
indoor space in the coldest winter months is assumed
to be not more than 25% in Zone 5, 20% in Zone
6, and 10% in Zone 7. For the precast concrete wall
with batt insulation, the relative humidity of the indoor
space in the coldest winter months is assumed to be
not more than 25% in Zone 5. The recommendations
were developed using these typical indoor relative humidities during winter. Indoor relative humidities greater than these during December, January, and February
have the potential to cause condensation within these
or any wall system not properly designed. Calculations
may be required when exterior sheathing is used on
the cold outdoor side since it may act as a vapor retarder on the cold side of the wall.
Fittings installed in outer walls, such as electrical

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

boxes without holes and conduits, should be completely sealed against moisture and air passage, and
they should be installed on the warm side insulation.
Also, high thermal conductance paths such as at connections inward from or near the colder surfaces may
cause condensation within the construction.
Warm Humid Climates (1A, 2A, 3A south of the
humid line). In these climates, the exterior surface
should have a lower vapor permeance than the interior
surface. For all three walls, the exterior concrete acts as
a semi-impermeable vapor retarder during the warm humid months. For the concrete sandwich panel wall, the
inside concrete wythe and painted gypsum wallboard, if
used, act as the semi-permeable vapor retarder during
the cool months. For the precast concrete wall with rigid
insulation, the insulation and painted gypsum wallboard
on the inside act as a semi-permeable vapor retarder during the cool months. For the precast concrete wall with
batt insulation, the painted gypsum wallboard acts as a
semi-permeable vapor retarder during the cool months.
Low permeance paints, vinyl wallpaper, or other materials that act as vapor retarders should not be placed on
the interior surface of the wall. Moisture from outdoors
often accumulates behind these materials when used in
these climates. Kraft paper is not recommended on the
insulation in these climates because it also prevents the
wall from drying.
In warm humid climates during rainy periods, exterior
walls can absorb large quantities of moisture that are
later driven inward by warm temperatures and solar effects. The concrete and rigid insulation (where provided) each have a moderately low permeance that helps
prevent this moisture from moving inward. Some exterior paints and finishes can also provide an adequate
level of resistance to moisture intrusion. The concrete
and rigid insulation should be continuous and sealed
to prevent the moisture from moving inward.
The operation of the cooling system is more important in warm and humid climates than any other climate. Since the latent load (that required to remove
moisture) is often greater than the sensible load (that
required to bring down the temperature), the system needs to be designed to remove the latent load
without cycling off because it has reached the desired
temperature set point. Oversized air-conditioners may
cycle off before the latent load is removed. Setting the
chilled water supply temperature too high will have
the same effect of not being able to remove the latent

5

load. Also, many people erroneously think that setting
the thermostat lower will remove moisture problems.
Low thermostat settings on hot humid days has the
opposite effect; they make surfaces colder and more
prone to condensation.
Warm Dry, Mixed, and Marine Climates (1B, 2B,
3A north of the humid line, 3B, 3C, 4). The need for
vapor retarders and low permeance materials is less in
these climates than in cold or warm humid climates.
Condensation can occur by the mechanisms discussed
for cold climates, but the duration of these conditions
is usually short enough that the materials subsequently
dry without problems if surfaces are semi-permeable
or semi-impermeable. The strategy for these climates is
to allow the wall system to dry either to the outside or
inside, or preferably, to both sides, since more damage
is caused by improperly placed vapor retarders than by
omitting one. The three precast concrete walls allow
this drying to either side. The exterior concrete wythe
acts as a semi-impermeable vapor retarder during the
warm months. For the concrete sandwich panel wall,
the inside concrete wythe and painted gypsum wallboard, if used, act as the semi-permeable vapor retarder during the cool months. For the precast concrete
wall with rigid insulation, the insulation and painted
gypsum wallboard on the inside act as a semi-permeable vapor retarder during the cool months. For the
precast concrete wall with batt insulation, the painted
gypsum wallboard acts as a semi-permeable vapor retarder during the cool months.
For the sandwich panel wall and the precast concrete
wall with XPS insulation, the relative humidity of the
indoor space in the coldest winter months is assumed
to be not more than 40% in Zone 4. For the precast
concrete wall with EPS insulation, the relative humidity of the indoor space in the coldest winter months is
assumed to be not more than 30% in Zones 4A and
4B and 35% in Zone 4C. For the precast concrete wall
with batt insulation, the relative humidity of the indoor
space in the coldest winter months is assumed to be
not more than 30% in Zone 4A and 35% in Zones 4B
and 4C. The recommendations were developed using
these typical indoor relative humidities during winter.
Indoor relative humidities greater than these during
December, January, and February have the potential to
cause condensation within these or any wall system
not properly designed.
These recommendations are for general use under normal building operating conditions.

ARCHITECTURAL PRECAST CONCRETE

|

429

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Backer Rod and Sealant

Moisture resistant, fire rated expandable
polyurethane foam in-place insulation
Exterior Surface
Spandrel

Spandrel

Backer Rod and Sealant (if access to interior is not
available, install from accessible side)

Painted Gypsum Wallboard with airspace, if desired
“A” Precast Sandwich Panel Wall
Not to scale
Backer Rod and Sealant
Spandrel

Exterior Surface
Spandrel

Backer Rod and Sealant (if access to interior is not
available, install from accessible side)

Painted Gypsum Wallboard with required insulation
“B” Precast Concrete with Rigid Insullation
Not to scale

Backer Rod and Sealant
Spandrel

Exterior Surface
Spandrel

Backer Rod and Sealant (if access to interior is not
available, install from accessible side)

Painted Gypsum Wallboard with required insulation

Fig. 5.3.21
Typical spandrel /column detail – Option A.

430

|

ARCHITECTURAL PRECAST CONCRETE

“C” Precast Concrete with Batt Insullation
Not to scale

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

Backer Rod and Sealant

5

Vapor Retarder rolled and pushed in joint
or foam insulation

Spandrel

Exterior Surface
Spandrel

Moisture resistant, fire rated expandable
polyurethane foam in-place insulation

Painted Gypsum Wallboard with airspace, if desired
“A” Precast Sandwich Panel Wall
Not to scale
Exterior Surface

Backer Rod and Sealant

Spandrel

Spandrel

Moisture resistant, fire rated expandable
polyurethane foam in-place insulation

Painted Gypsum Wallboard with required insulation
“B” Precast Concrete with Rigid Insulation
Not to scale

Backer Rod and Sealant
Spandrel

Exterior Surface
Spandrel

Vapor Retarder rolled and pushed in joint
or foam insulation
Painted Gypsum Wallboard with required insulation

Fig. 5.3.22
Typical spandrel /column detail – Option B.

“C” Precast Concrete with Batt Insulation
Not to scale

ARCHITECTURAL PRECAST CONCRETE

|

431

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Rigid Insulation (if not relying on composite action)*
IECC Zone
Min. R-Value
1
None
2, 3
R5
4, 5, 6
R10
R15
7, 8

Topping Slab
OFFICE

*NOTE: Add topping slab below insulation
for composite action

(1) Layer 5/8-in. type ‘X’ M.R. gypsum
board on grid system

Precast

Rigid Insulation
(if not relying on composite action)*
IECC Zone
Min. R-Value
1
None
2, 3
R5
4, 5, 6
R10
7, 8
R15

Precast Concrete Spandrel
PARKING

Topping Slab
OFFICE

“A” Precast Sandwich Panel Wall
Not to scale

Precast

(1) Layer 5/8-in. type ‘X’ M.R. gypsum
board on grid system

Precast Concrete Spandrel
PARKING

Rigid Insulation
(if not relying on composite action)*
IECC Zone
Min. R-Value
1
None
2, 3
R5
4, 5, 6
R10
R15
7, 8

Topping Slab
OFFICE

Precast

“B” Precast Concrete with Rigid Insulation
Not to scale

(1) Layer 5/8-in. type ‘X’ M.R. gypsum
board on grid system

Precast Concrete Spandrel
PARKING

(See Fig. 5.3.25 for Spandrel/DT insulation}

“C” Precast Concrete with Batt Insulation
Not to scale

432

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.23 Typical floor detail – rigid insulation.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

5

Topping Slab
Precast

OFFICE

(1) Layer 5/8-in. type ‘X’ M.R. gypsum board on grid system

Precast
Concrete
Spandrel

Spray-on or Batt Insulation
IECC Zone
Min. R-Value of added insulation
1
None
2, 3, 4, 5, 6
R11
R15
7, 8

PARKING
Topping Slab
Precast

OFFICE

“A” Precast Sandwich Panel Wall
Not to scale
(1) Layer 5/8-in. type ‘X’ M.R. gypsum board on
grid system
Precast
Concrete
Spandrel

Spray-on or Batt Insulation
IECC Zone
Min. R-Value of added insulation
1
None
2, 3, 4, 5, 6
R11
R15
7, 8

PARKING

OFFICE

Topping Slab
Precast

“B” Precast Concrete with Rigid Insulation
Not to scale

(1) Layer 5/8-in. type ‘X’ M.R. gypsum board on grid system

Precast
Concrete
Spandrel

Spray-on or Batt Insulation
IECC Zone
Min. R-Value of added insulation
1
None
2, 3, 4, 5, 6
R11
R15
7, 8

PARKING

(See Fig. 5.3.25 for Spandrel/DT insulation}

“C” Precast Concrete with Batt Insulation
Not to scale

Fig. 5.3.24 Typical floor detail – alternate batt insulation.

ARCHITECTURAL PRECAST CONCRETE

|

433

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Building Felt
Topping Slab

Tooled &
JT. Sealant

Furring and Gypsum Wallboard, as desired
“A” Precast Sandwich Panel Wall
Not to scale

Moisture resistant, expandable
polyurethane foam in-place insulation
Topping Slab

Tooled &
JT. Sealant

Moisture resistant, expandable
polyurethane foam in-place insulation
Topping Slab

Rigid Installation, Furring and Gypsum
Wallboard as required
“B” Precast Concrete with Rigid Insulation
Not to scale

Tooled &
JT. Sealant

Batt Installation, Metal Framing and
Gypsum Wallboard as required

“C” Precast Concrete with Batt Insulation
Not to scale

434

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.25 Typical non-loadbearing spandrel/DT detail.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

5

Backer Rod and Sealant
Glass

Flashing or Slope Shelf Outward

Backer Rod
and Sealant

5 8

/ -in. thick moisture resistant gypsum wallboard

Insulation

R2 rigid insulation (min.), wood or composite wood block in IECC Zones 4 – 7

Precast
Concrete
Panel

‘Z’ furring or metal framing to create 1-in. minimum air space

/ -in. thick gypsum wallboard painted (optional)

5 8

“A” Precast Sandwich Panel Wall
Not to scale

Glass

Flashing or Slope Shelf Outward

Backer Rod
and Sealant

Backer Rod and Sealant
5 8
/ -in. thick moisture resistant
gypsum wallboard or other sill material

R2 rigid insulation (min.), wood or
composite wood block in IECC Zones 4 – 7
‘Z’ furring or metal framing to create
1-in. minimum air space

Precast
Concrete
Panel

/ -in. thick gypsum wallboard painted

5 8

Rigid or semi-rigid insulation, continuous
and taped at joints, as required
“B” Precast Concrete with Rigid Insulation
Not to scale

Glass

Flashing or Slope Shelf Outward
Backer Rod and Sealant

Backer Rod
and Sealant

/ -in. thick moisture resistant gypsum wallboard or other sill material

5 8

R2 rigid insulation (min.), wood or composite wood block in IECC Zones 4 and 5
1-in. minimum air space
Precast
Concrete
Panel

/ -in. thick gypsum wallboard painted

5 8

Batt insulation in metal framing as required
Only applicable for IECC Zones 1 – 5
“C” Precast Concrete with Batt Insulation
Not to scale

Fig. 5.3.26 Typical window sill detail.

ARCHITECTURAL PRECAST CONCRETE

|

435

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Precast
Concrete
Panel
Acoustical Ceiling

‘Z’ furring or metal framing to create 1-in. minimum air space

Insulation

R2 rigid insulation (min.), wood or composite wood block in IECC Zones 4 – 7

Drip Edge

/ -in. thick gypsum wallboard, painted (optional)

5 8

“A” Precast Sandwich Panel Wall
Not to scale

‘Z’ furring or metal framing to
create 1-in. minimum air space
Rigid insulation, continuous and
taped at joints, as required

Precast
Concrete
Panel

Acoustical Ceiling

R2 rigid insulation (min.), wood
or composite wood block in IECC
Zones 4 – 7
Drip Edge

/ -in. thick gypsum wallboard, painted

5 8

“B” Precast Concrete with Rigid Insulation
Not to scale
1-in. minimum air space

Precast
Concrete
Panel

Batt insulation in metal framing as required

Acoustical Ceiling

R2 rigid insulation (min.), wood or composite wood block in IECC Zones 4 and 5
Drip Edge
/ -in. thick gypsum wallboard painted

5 8

Only applicable for IECC Zones 1 – 5
“C” Precast Concrete with Batt Insulation
Not to scale

436

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.27 Typical window head detail.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

5

Building Felt
Tooled & JT. Sealant

DT

Topping

Spot corbel or ledge

Tooled & JT. Sealant
At Web, Flange and Floor:
Fill space between precast
web, flange, floor and exterior
wall with mosture resistant,
foam in-place expandable
polyurethane insulation

Topping

DT

“A” Precast Sandwich Panel Wall
Not to scale

BEtWEEN Web and CORBEL:
Extend insulation and gypsum
wallboard to bottom of precast
concrete flange

Spot corbel or ledge

Wrap corbel with required rigid insulation
and 5/8-in. thick gypsum wallboard
“B” Precast Concrete with Rigid Insulation
Not to scale
Tooled & JT. Sealant
Topping

DT

At Web, Flange and Floor:
Fill space between precast
web, flange, floor and exterior
wall with mosture resistant,
foam in-place expandable
polyurethane insulation

BEtWEEN Web and CORBEL:
Extend insulation and gypsum
wallboard to bottom of precast
concrete flange

For Sections B & C – corbel insulation
IECC Zone
Min. R-Value of Added Insulation
1, 2
None
3B
None
None
3A (Below “Warm Humid Line”)
3A (Above “Warm Humid Line”)
R7.5
3C, 4A, 4B
R7.5
4C, 5, 6*
R10
7
R12.5
*Not Applicable for Section C

Spot corbel or ledge
Wrap corbel with required rigid insulation
and 5/8-in. thick gypsum wallboard

“C” Precast Concrete with Batt Insulation
Not to scale

Fig. 5.3.28
Typical loadbearing spandel with corbel/DT detail.

ARCHITECTURAL PRECAST CONCRETE

|

437

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

Spot
corbel
or ledge

DT

Topping

Varies

Building Felt
Tooled & JT. Sealant

Ledge
“A” Precast Sandwich Panel Wall
Not to scale

Topping

Spot
corbel
or ledge

BEtWEEN Web and CORBEL:
Extend insulation and gypsum
wallboard to bottom of precast
concrete flange

DT

At Web, Flange and Floor:
Fill space between precast
web, flange, floor and exterior
wall with mosture resistant,
foam in-place expandable
polyurethane insulation

Varies

Tooled & JT. Sealant

Wrap corbel with required rigid insulation
and 5/8-in. thick gypsum wallboard
Ledge

“B” Precast Concrete with Rigid Insulation
Not to scale

Topping

Spot
corbel
or ledge
BEtWEEN Web and CORBEL:
Extend insulation and gypsum
wallboard to bottom of precast
concrete flange

|

For Sections B & C – corbel insulation
IECC Zone
Min. R-Value of Added Insulation
1, 2
None
3B
None
None
3A (Below “Warm Humid Line”)
3A (Above “Warm Humid Line”)
R7.5
3C, 4A, 4B
R7.5
4C, 5
R10

Wrap corbel with required rigid insulation
and 5/8-in. thick gypsum wallboard
Ledge

“C” Precast Concrete with Batt Insulation
Not to scale

438

DT

At Web, Flange and Floor:
Fill space between precast
web, flange, floor and exterior
wall with mosture resistant,
foam in-place expandable
polyurethane insulation

Varies

Tooled & JT. Sealant

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.29
Typical loadbearing spandrel with corbel /dapped DT detail.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

5

Building Felt
Tooled & JT. Sealant
Concrete topping slab

Pocketed Precast Spandrel

DT

Precast

“A” Precast Sandwich Panel Wall
Not to scale
Tooled & JT. Sealant
Fill space between precast web, floor and exterior wall with
mosture resistant, foam in-place expandable polyurethane
insulation
Concrete topping slab

Pocketed Precast Spandrel

Fill pocket with required
insulation in IECC Zones 3 – 7

Extend required insulation and
gypsum wallboard to bottom of
precast concrete flange
DT

Precast

“B” Precast Concrete with Rigid Insulation
Not to scale
Tooled & JT. Sealant
Fill space between precast web, floor and exterior wall with
mosture resistant, foam in-place expandable polyurethane
insulation

Pocketed Precast Spandrel

Concrete topping slab

Fill pocket with required
insulation in IECC Zones 3, 4, and 5

Extend required insulation and
gypsum wallboard to bottom of
precast concrete flange
DT

“C” P
 recast Concrete with Batt Insulation
Not to scale

Precast

Fig. 5.3.30
Typical pocketed loadbearing spandrel /DT detail.

ARCHITECTURAL PRECAST CONCRETE

|

439

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

EPDM Membrane, Follow National
Roofing Contractors Association
Recommended Details

EPDM Membrane, Follow National
Roofing Contractors Association
Recommended Details
Precast Concrete
Panel

Roof Membrane on Tapered
Rigid Insulation

Tooled &
Sealed Joint

Roof Membrane on Tapered
Rigid Insulation

Precast Concrete
Panel

DT

DT

Tooled &
Sealed Joint

OR

“A” Precast Sandwich Panel Wall
Not to scale

“A” Precast Sandwich Panel Wall
Not to scale

EPDM Membrane, Follow National Roofing
Contractors Association Recommended Details
Roof Membrane on Tapered Rigid Insulation

Precast Concrete
Panel

DT

Tooled &
Sealed Joint

Mosture resistant, fire rated foam in-place
expandable polyurethane insulation

Rigid Insulation, Continuous & Taped
at Joints, as required
“B” Precast Concrete with Rigid Insulation
Not to scale

EPDM Membrane, Follow National Roofing Contractors
Association Recommended Details
Precast Concrete
Panel

Roof Membrane on Tapered Rigid Insulation

Tooled &
Sealed Joint

DT

Mosture resistant,
fire rated foam
in-place expandable
polyurethane
insulation

Batt Insulation in Metal Framing, as required
“C” Precast Concrete with Batt Insulation
Not to scale

440

|

ARCHITECTURAL PRECAST CONCRETE

Fig. 5.3.31 Typical roof parapet detail.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

Special Applications and Building Type. Special
precautions are required for buildings with high indoor
humidities or spaces with sensitive electronic equipment or artifacts. These include swimming pools, ice
rinks, cold storage, computer rooms, libraries, hospitals, nursing homes, museums, and some manufacturing facilities. Low permeance vapor retarders are often
needed to separate indoor swimming pools or other
special applications from the rest of the building.
Details. Figures 5.3.21 through 31 provide conceptual details on how to construct the precast concrete
system to achieve energy savings while providing an air
barrier and reducing the potential for moisture problems. The recommendations and details presented
are based on specific analyses, engineering judgment,
and best available practices at the time of publication. Performance testing of the details has not been
performed. Detail drawings are provided in order to
assist competent professionals in the detailing of the
building insulation envelope. Reinforcing designations,
structural connections, wythe thickness, and insulation
indicated in drawings are to be used for reference only
and are not intended to substitute for project specific
judgment.
Water Leakage. The exterior surface of the precast
concrete system acts as a weather barrier to prevent
rain and snow from entering the building. As shown
in Figs. 5.3.21 and 22, joints in the precast concrete
generally have either two layers of sealant, or sealant
and a secondary method of defense against water
penetration. Joints around windows, doors, and other
penetrations through the precast concrete building are
designed with a primary and secondary method of defense against rainwater penetration.
Floor Systems. The provided details are for a double
tee floor system. Details for hollow core floor systems
will be similar, including insulation requirements. The
main concept is to separate the floor slabs from the exterior concrete by insulation to reduce thermal bridges. This will reduce energy losses and the potential for
condensation and moisture problems.
Figures 5.3.23 and 24 present two options for insulating floors above unconditioned spaces such as parking structures. In these cases the concrete floor acts
as a semi-impermeable vapor retarder. Figure 5.3.23
with rigid insulation is preferable. If spray-on or batt
insulation are used as shown in Fig. 5.3.24, it should
be wrapped around the precast concrete stems.

5

Need for analysis. In any building with additional
sources of moisture, such as from swimming pools,
industrial processes, or storage of moist items, a moisture analysis of the walls and roofs for actual conditions
should be performed. For instance, hospitals in cold
climates often maintain RH levels at 50%, as opposed
to the 20 to 40% RH in most buildings during winter
in these climates. This higher RH can cause moisture
problems if the building envelope is not properly designed. An analysis is also advisable for very cold, cold,
mixed, or cool marine climates or other climates where
experience is not available to indicate how a wall will
perform. It is important to determine whether and
where the temperature within the envelope system will
fall below the dew point temperature. Accurate analyses take into account moisture absorption of materials
as well as moisture movement through walls.
ASTM publishes an excellent book on moisture models.17 These models predict moisture and temperature
conditions in wall and roof assemblies for particular climate and indoor design conditions. The models utilize
mathematical solutions to moisture and heat transfer
mechanisms. Some predict moisture transfer by air
movement and liquid water flow as well as vapor diffusion. Some can model the changes in material properties such as permeance and sorption with moisture
content. Use of these models requires knowledge of
building physics, material properties, and the model
limitations.
Historically a simplified method known as the dew
point method has been used to identify potential
condensation problems. This is a simplified steadystate analysis that has many limitations. If used with
worst case conditions that only take place a few days
a year, it will identify condensation that may not be
a problem due to the ability of the materials to absorb the moisture or for the system to dry within a
few days. For this reason, monthly averages are generally used. Since it considers only steady-state conditions, it is not exact. The vapor diffusion properties
of materials often vary with moisture content, which
are not considered in the dew point method. Also,
it is often frequently misused in identifying where
and how much condensation occurs. However, the
dew point method is a good indicator of the potential for moisture problems. The ASHRAE Handbook
of Fundamentals and ASTM C 755 provide excellent
17 Treschsel, Heinz, Moisture Analysis and Condensation Control in Building
Envelopes, Publication No. MNL 40, ASTM, West Conshohocken, PA,
2001. www.ASTM.org

ARCHITECTURAL PRECAST CONCRETE

|

441

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.4 Condensation within walls and use of vapor retarders

each side of the wall.

Example No. 5.3.6 Condensation within exterior wall.

Step 1. The vapor pressures of the indoor and outdoor air may be determined from saturated vapor pressures listed in Table 5.3.15 and the assumed temperatures and relative humidities. The actual vapor pressure
is the saturated vapor pressure at the given temperature times the relative humidity:

/ in. Gypsum Wallboard Light Gray
Concrete Finish

1 2

6 in. Concrete

INTERIOR
T = 70F
R.H. = 30%

2 in. Insulation

1 Coat Flat Oil Paint
on 1 Coat of Primer
(Typical)

EXTERIOR
Tmax = 95F
Tmin = 0F
R.H. = 80%

The selection of appropriate outside air temperatures
requires considerable judgment. The effects of heat stor-

81/2 in.

Table 5.3.15 Vapor Pressures.

descriptions and examples of the dew point method.
An example also follows.

Temp., RH,
ºF
%

An analysis will be performed to determine whether
condensation will form within the wall for the temperature and relative humidity conditions indicated on
Thermal Resistance and Temperature of Insulated Wall.

R-Value
Winter

Temp.
Temp.,
Difference,
ºF
ºF
70

Surface,
inside

A.

0.68

0.45

Surface,
outside

|

0.222

Outdoor

0

80

0.038

0.030

SVP, in.
Hg

M or
μ/n,
perm

zn, rep

Δpn,
in.
Hg

0.739
Primer and
paint, 1 coat

0.622

0.560

B.

2

Gypsum
wallboard, 1/2 in.

8.00

58

C.

0

0

0.5

0.033
0.189

38

0.38

EPS insulation
(1.25 pcf), 2 in.

0.046

0.027

0.002

3

0.187
4/2
=2

0.5

0.033

0.046
D.

Concrete, 6 in.
(145 pcf)

0.154
3.2/6
= 0.53

1.87

0.124

0.040

0.17

Total

9.68

U = 1/R

0.10

0.040

E. Surface, outside

0.030
0

0

0.038

1

Total

70

ARCHITECTURAL PRECAST CONCRETE

0.038

Pa,
in.
Hg
0.222

0.560

0

442

0.739

0.622

1
E.

30

SVP,
in.
Hg
0.739

3

4
Concrete,
D.
6 in.
(145 pcf)

70

A. Surface, inside

62
EPS
insulation
C.
(1.25 pcf),
2 in.

Indoor

Vapor Resistance and Vapor Pressure for Continuity.

5
65

Gypsum
B. wallboard,
1
/2 in.

Vapor Pressure Actual Vapor
at Saturation,
Pressure,
in. Hg
in. Hg

0.030
2.90

0.192

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders

age in materials must be recognized, as must the fact that
wall or roof surface temperatures can be higher than air
temperature because of solar radiation, and colder than
air temperature because of clear sky radiation. These temperature modifications vary with the color, texture, thickness, weight and orientation of the surface materials and
with the intensity of the radiation. Generally the average
January temperatures without solar effects and the average July temperatures with solar effects are recommended
for determining the potential for condensation. The effect
of solar radiation and humid outdoor conditions alter the
dew point. Most building veneer systems are not waterproof and absorb moisture. When this moisture is heated
by the sun, the vapor pressure in the veneer increases and
drives the moisture inward.
Step 2. Determine the thermal resistance of the wall
and temperatures within the wall using Eqs. 5.3.1 and
5.3.2 as in Example 5.3.1 (page 407) and 5.3.5 (page
421):
Thermal bridges are not considered in this example
and would need to be analyzed separately.
The temperature existing at any point in a wall under
any given exterior and interior temperature conditions

Fig. 5.3.32 Thermal and water vapor gradients for extreme
winter conditions.
65
Temperature, F

60

Step 3. The saturated vapor pressures at various surfaces and interfaces within the wall section may be
obtained from temperatures determined in Step 2 and
Table 5.3.14, page 425.
These saturated vapor pressures (SVP or Ps) are plotted in Fig. 5.3.32 to form the SVP gradient, Ps, through
the wall section.
Step 4. To check the location where condensation is
likely to take place, the vapor pressure gradient necessary for vapor transfer continuity, Pc, is plotted as
shown in Fig. 5.3.32. The vapor pressure gradient, Pc,
is obtained by a calculation procedure similar to that
used to determine the temperature gradient, described
in Step 2. It is based upon the total vapor pressure drop
(0.222 – 0.030 = 0.192 in. Hg) and the respective vapor permeances of the different components of the
wall from Table 5.3.12.

Δpn

20
4

1

-20
0.7
0.6

0.56

= vapor pressure gradient or drop through
material “n”, in. of mercury

zn

= vapor pressure resistance of material “n”, rep
(rep = 1/perm)

zwall

= vapor pressure resistance of wall, rep

and

0.4

zn

Ps

0.3

0.1

Equation No. 5.3.6

Δpwall = vapor pressure gradient or drop through wall,
in. of mercury

0.62

0.5

0.2

= zn(Δpwall) / zwall

where:

40

0

Vapor Pressure, In. Hg

is of great significance in designing problem-free building enclosures. An ability to calculate the thermal gradient permits the designer to forecast the magnitude
of the movements caused by external temperature
changes, to predict the location of condensation and
freezing planes in the wall, and to assess the suitability
of any construction. The temperature gradient will not,
in itself, give the designer all the information required
to select and assemble building components, but it is
an essential first step.

Δpn

62

5

0.22

= 1/M or n/μ

Equation No. 5.3.7

where:
0.19

0.15

0.06

0.0
Condensation Surface

Pc

0.04
0.03

M

= vapor permeance, perms

n

= thickness of material, in.

µ	

= vapor permeability, perm·in.

Continuous vapor flow conditions are preserved pro-

ARCHITECTURAL PRECAST CONCRETE

|

443

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.4 Condensation within walls and use of vapor retarders / 5.3.6.5 Air infiltration, exfiltration, and air barriers

vided the actual vapor pressure, Pa, does not exceed the
saturation vapor pressure, Ps. If Pa does exceed or cross
Ps, condensation will occur. In this case, Pa exceeds Ps
in between the insulation and concrete layers and condensation is expected to occur here. For discontinuous vapor flow (when condensation occurs), the vapor
flow to and away from the condensation surface must
be recalculated. The difference will be equal to the
condensation rate. The vapor flow to or from a point
is equal to the actual vapor pressure difference divided
by the vapor resistance to or from that point.
Reducing Pa so that it is less than Ps can be achieved
either by:
1. Changing the various vapor flow resistances to reduce the values of Pa. For example, add a vapor
retarder on the side of the wall with the higher vapor pressure (warm side) or use an insulation with
a lower vapor permeance.
2. Changing the various thermal resistances of the
wall components to raise the temperature. This
will raise the values of Ps.
3. A combination of two of the above items.

5.3.6.5 A
 ir infiltration, exfiltration,
and air barriers
Infiltration and exfiltration are air leakage into and
out of a building respectively, through cracks or joints between infill components and structural elements, interstices around windows and doors, between the sill plate
and foundation, through floors and walls, at the top and
bottom of walls, and at openings for building services
such as plumbing. Approximately 5 to 20% of air leakage occurs at doors and windows, and 20 to 50% occurs through walls. Infiltration and exfiltration are often
a major source of energy loss in buildings. Exfiltrating air
carries away heating and cooling energy, while infiltrating air may bring in moisture and pollution as well as
reduce the effectiveness of a rain screen wall system.
Moisture can move into or across a wall assembly by
means of vapor diffusion and air movement. Diffusion
is a slow, controlled process driven only by vapor pressure differentials, and rarely causes any significant
moisture accumulation. Air migration occurs from air
pressure differentials independent of moisture pressure differentials. If air, especially exfiltrating, warm,
humid air, can leak into the enclosure, then this will
be the major source of moisture. Condensation due to

444

|

ARCHITECTURAL PRECAST CONCRETE

air movement is usually much greater than that due
to vapor diffusion for most buildings. However, when
air leakage is controlled or avoided, the contribution
from vapor diffusion can still be significant. In a well
designed wall, attention must therefore be paid to the
control of both air flow and vapor diffusion.
An air barrier and vapor retarder are both needed in
a properly designed building envelope, and in many instances a single material can be used to provide both of
these as well as other functions. The principal function
of the air barrier is to stop the outside air from entering
the building through the walls, windows, or roof, and
inside air from exfiltrating through the building envelope
to the outside. This applies whether the air is humid or
dry, since air leakage can result in problems other than
the deposition of moisture in cavities.
Uncontrolled air (and its associated water vapor), exfiltration in cold climates and infiltration in hot, humid
climates can wreak havoc on the structure, causing
corrosion and structural damage, mold and bacterial
growth, and energy loss. It can also create HVAC problems by disrupting indoor air pressure relationships and
degrading indoor air quality (IAQ), which can lead to
health problems for sensitive individuals.
Atmospheric air pressure differences between the inside and outside of a building envelope exist because
of the action of wind, the density difference between
outside cold heavy air and inside warm light air creating
a “stack effect”, and the operation of equipment such
as fans. The pressure differences will tend to equalize, and the air will flow through holes or cracks in
the building envelope carrying with it the water vapor
it contains. A thorough analysis of air leakage is very
complex, involving many parameters, including wall
construction, building height and orientation.
Air barriers (sometimes called air retarders) will reduce infiltration and exfiltration. They reduce the potential for moisture problems due to moist air migrating into a building. This moisture can be warm humid
air from outside during the summer or warm conditioned air from inside in the winter.
An air barrier is required to have a leakage rate less
than 0.06 cfm/ft2 at a differential pressure of 0.3 in.
H2O (1.57 psf) according to ASTM E1677, “Standard
Specification for an Air Retarder (AR) or Material or
System for Low-Rise Frame Walls.” This value however is considered high for buildings in Canada where
a value of 0.004 cfm/ft2 at 0.3 in. H2O is sometimes

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.5 Air infiltration, exfiltration, and air barriers

Table 5.3.16 – Measured air leakage for selected building materials1.

Material

Average leakage,
cfm/ft2 of surface
at 0.3 in. H2O

Solid precast concrete wall

No measurable
leakage

Aluminum foil vapor barrier

No measurable
leakage

6 mil polyethylene

No measurable
leakage

Extruded polystyrene
insulation

No measurable
leakage

Closed cell foam insulation

0.0002

3 in. polyicynene

0.001

1

/2 in. fiberboard sheathing

Breather type building
membranes

0.31
0.0022 – 0.71

Uncoated brick wall

0.31

Uncoated concrete block

0.41

1 in. expanded polystyrene

0.93

National Research Council, Canada www.nrc-cnrc.gc.ca

1

required. This is the maximum air leakage for a total assembled air barrier system (total wall system or
main areas plus joints) when tested according to ASTM
E2178, “Standard Test Method for Air Permeance of
Building Materials.”
Materials such as precast concrete panels, polyethylene, gypsum board, metal sheeting or glass qualify
as air barriers since they are low air-permeable materials when joints are properly sealed; concrete block,
acoustic insulation, open cell polystyrene insulation, or
fiberboard are not. Air permeances of selected materials are presented in Table 5.3.16.
Materials and the method of assembly chosen to build
an air barrier must meet several requirements if they are
to perform the air leakage control function successfully.
1. C
 ontinuous. The air barrier must be continuous
throughout the building envelope. For example,
the low air permeability materials of the wall must
be continuous with the low air barrier materials of
the roof (e.g., the roofing membrane) and connected to the air barrier material of the window
frame. All of the air barrier components should be

5

sealed together so there are no gaps in the envelope airtightness. Where interior finishing (drywall) serves as the air barrier, if it is not finished
or continuous above suspended ceilings or behind
convector cabinets, there will be large gaps in the
air barrier system’s continuity. Connection should
be made between:
a. Foundation and walls.
b. Walls and windows or doors.
c. Different wall systems, such as brick and precast concrete, or curtain wall and precast concrete, and corners.
d. Joints in gypsum wallboard and precast concrete panels.
e. Walls and roof.
f. Walls, floors and roof at construction, control,
and expansion joints. The interior air barrier
above a dropped ceiling needs to be connected
to the underside of the above floor.
g. Walls, floors, and roof to utility, pipe, and duct
penetrations.
2. Load Capacity. Each membrane or assembly of materials intended to support a differential air pressure
load must be designed and constructed to carry that
load, inward or outward. This load is the combined
wind, stack, and fan pressures on the building envelope. If the air barrier system is made of flexible materials, then it must be supported on both sides by
materials capable of resisting the peak air pressure
loads; or it must be made of self-supporting materials, such as board products adequately fastened to
the structure. The air barrier should be designed so
that adjacent materials are not displaced under differential air pressures. Tape and sealant must also
resist these pressures and have long-term resistance.
Concrete is the ideal material for an air barrier because of its durability and strength in resisting these
loads. Sealant between panels and at joints must be
designed to resist these loads.
3. Joints. The air barrier of each assembly should be
joined to air barriers of adjacent assemblies in a
manner allowing for the relative movement of the
assemblies and components due to thermal and
moisture variation, creep, and structural deflection.
These joints in the air barrier and joints at penetrations of the air barrier system should be of low air
permeability materials.

ARCHITECTURAL PRECAST CONCRETE

|

445

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.6.5 Air infiltration, exfiltration, and air barriers

4. Durable. The air barrier assembly must be durable
in the same sense that the building is durable, and
be made of materials that are known to have a
long service life or be positioned so that they may
be serviced from time to time.
5. Vapor Permeance. When a vapor retarder is used
on the inside of insulation in a cold or mixed climate [see vapor retarder section, page 423], an air
barrier used on the outside should be permeable
to water vapor. If both the inside vapor retarder
and the outside air barrier are not permeable, then
a “double vapor retarder” condition is created.
Moisture that gets between the two through rain
penetration or leakage through joints will not be
able to readily evaporate and disperse to the interior or exterior. Vapor permeability allows moisture
behind the air barrier to exit the wall by vapor diffusion to the outside. According to ASTM E1677:
“In a moderate to cold climate the opaque wall
must either be permeable to water vapor, or when
the permeance of the materials on the exterior is
less than 1 perm it may be beneficial to insulate
on the outside. When the exterior is permeable,
moisture vapor from the opaque wall can escape
to the outdoors without accumulating in the wall.
When the exterior is insulated, the temperature of
the opaque wall is increased to minimize wall moisture accumulation. Designers should evaluate the
amount of insulation necessary to keep condensation from forming in the wall assembly when the
air barrier is rated as a vapor retarder less than 1
perm and exterior applied.”
Building Pressure. In warm humid climates, a positive building pressure will help prevent the infiltration of humid air. In cold climates the building pressure should be neither strongly positive or negative.
A strong negative pressure could pull in combustion
products from street traffic. A strong positive pressure
could drive moisture into the building walls and other
elements.
Adequate Ventilation. Because concrete buildings
have less air leakage, heating and cooling systems
should have adequate air intake systems to provide
fresh air in buildings. This is more critical in concrete
than steel frame buildings because there is less air
leakage. Without an adequate intake source, concrete
buildings are under negative pressure, potentially resulting in poor indoor air quality. In all cases, guidelines

446

|

ARCHITECTURAL PRECAST CONCRETE

of ANSI/ASHRAE Standard 6218 should be followed for
proper ventilation of indoor air.
Application. The location of the air barrier in the
wall system is dependent on the wall construction and
climate. Precast concrete as a material acts as an air
barrier and has a negligible air leakage and infiltration
rate. A properly designed and constructed precast concrete building will save energy due to this low infiltration. This requires the air barrier be continuous by
sealing joints between precast concrete panels, openings at connectors, around door and window frames,
and at penetrations. The building envelope should provide continuous resistance to air flow through joints at
floors, ceilings, and roof. Gypsum wallboard can act
as an air barrier if the floor/wallboard and ceiling/wallboard joints are tightly fitted and sealed with a joint
sealant.
Air barrier membranes and building wraps such as
Tyvek® are being used more frequently in new construction. They are not required in precast concrete
buildings because the concrete acts as an air barrier
and has a lower air permeance than many of the available membranes and wraps (see “breather type membranes” in Table 5.3.16).
In cold climates (Zones 5, 6, and 7), it is strongly recommended that the visible interior surface
of a building envelope be installed and treated as
the primary air barrier and vapor retarder. A concrete panel with the concrete on the indoor surface
generally serves this dual function as air barrier and
vapor retarder. Where floors and cross walls are of
solid concrete, it is necessary to seal only the joints, as
floors and walls themselves do not constitute air paths.
Where hollow partitions, such as steel studs, are used,
the interior finish of the envelope can be made into the
continuous air barrier. Where this is impractical, polyethylene film should be installed across these junctions
and later sealed to the interior finish material. Where
it is impractical to use a concrete panel system as the
continuous air barrier system, an interior finish of gypsum wallboard, or plaster, painted with two coats of
vapor retarding paint, will provide a satisfactory air
barrier/vapor retarder in many instances.
While it is preferable that the air barrier system be
placed on the warm indoor side of an insulated assem18 ASHRAE Standard 62-2004, "Ventilation for Acceptable Indoor
Air Quality," American Society of Heating, Refrigerating, and AirConditioning Engineers, Atlanta, www.ASHRAE.org

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.6.5 Air infiltration, exfiltration, and air barriers / 5.3.7 Application of Insulation

bly, where thermal stresses will be at a minimum, it is
not an essential requirement. (This does not necessarily
mean on the inside surface of the wall.) The position
of the air barrier in a wall is more a matter of suitable
construction practice and the type of materials to be
used. However, if an air barrier membrane is used and
is positioned on the outside of the insulation, consideration must be given to its water vapor permeability, as
discussed in Item No. 5. One rule of thumb is to choose
an air barrier material on the outside that is 10 to 20
times more permeable to water vapor diffusion than
the vapor retarder material on the inside of the wall.
In warm and humid climates (Zones 1A, 2A, and 3A),
an air barrier (or low air-permeance materials properly
sealed) on the outside of the wall works well because
it helps prevent the infiltration of the warm humid air.
An architectural precast concrete panel with appropriate joint sealant will serve as an air barrier in this climate. Exterior surfaces should be less permeable than
inside surfaces, once again, to help reduce the amount
of moisture entering the walls. Note that this is the opposite of what is recommended for cold climates.
In mixed, dry warm, and cool marine climates
(see Table 5.3.11), an air barrier (or properly sealed
low air-permeance materials) is recommended. An
architectural precast concrete panel with appropriate
joint sealant will serve as an air barrier in this climate.

5.3.6.6 C onsiderations at windows
The principal potential moisture problems with windows are the following:
1. Poor sealing of the wall air barrier and vapor retarder at window joints with the wall.
2. Penetration of rainwater into the wall construction
beneath the windows.
3. Condensation of moisture or frost formation on
the inside of windows in cold weather and subsequent drainage of the water onto the sill and into
the wall construction.
4. Excessive leakage of warm moist air into the building in summer weather that adds to the air conditioning load.
Air barriers and vapor retarders must be carefully
sealed at window openings to prevent air leakage
into wall construction at the window frames. Likewise
the design of window sills and the sealant techniques
must be such that rainwater drainage is diverted to the

5

outside without wetting the insulated construction beneath the windows. This requires that thermal insulation be held away from the collecting surface so moisture can proceed down to collection systems without
wetting the insulation. Impaling pins allow this to be
accomplished easily, and they are available with shoulders holding back-up discs and insulation away from
the panel.
Double and triple glazed windows should be used in
Climate Zones 4, 5, 6, and 7 where there are extended
periods of cold weather to reduce surface condensation and drainage. An indoor relative humidity of 40%
can be maintained without excessive condensation
on double-glazed windows for outside temperatures
down to 15ºF. At colder temperatures, indoor RH levels are generally lower and the potential for condensation will generally be lower. Windows with argon fill
allow for colder temperatures before condensate accumulates. The ASHRAE Handbook of Fundamentals
provides more guidance on condensation. The drainage of window condensation should not be allowed to
remain on the window sills or to run down the inside
walls. Windows in hospitals and swimming pool areas
are exposed to higher than average indoor RH levels in
cold climates and must be carefully designed to prevent condensation.
Excessive window leakage can be avoided by specifying the maximum acceptable leakage observed when
windows are tested in accordance National Fenestration
Rating Council (NFRC www.NFRC.org) Test Method
400. Air leakage should not exceed 0.4 cfm/ft2. These
values are available from the manufacturer.

5.3.7 Application of Insulation
Where wall insulation is required in a building, it may
be applied to the precast concrete panel (normally to
the interior surface) or it may be fully incorporated in
the precast concrete panel, resulting in a sandwich
wall panel.
There are several methods to apply insulation to large,
flat precast concrete surfaces:
1. Supplementary framing (e.g. steel studs) can be
added to provide cavities for the installation of
batts or rigid insulation and to support subsequent
components of the assembly. There should be an
air space between the framing and the panels to
minimize thermal bridging. Additionally, batts and

ARCHITECTURAL PRECAST CONCRETE

|

447

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.7 Application of Insulation

other moisture sensitive materials should never be
in contact with concrete, especially concrete that
is subjected to wetting by rain or other sources of
moisture.
2. Rigid insulation can be fastened to concrete surfaces with adhesives, by impaling it on adhered
pins (“stick clips”), and with various types of furring and mechanical fasteners.
Adhesives: This is the most obvious method of fastening anything to a large flat surface and there are a
number of adhesives available for this use. Selection of
the proper adhesive is important. It should be compatible with the type of insulation being used. The vehicles
or thinners in some adhesives will attack foam plastic insulation. Also, some protein-based adhesives can
provide nutrition for fungi and other micro-organisms
unless they have preservatives included in their makeup.
The adhesive should not be applied in daubs. The use
of daubs of adhesive creates an air space between the
surface and the insulation. If the insulation is on the inner surface of the assembly, warm moist air circulating
in this space will cause condensation. If the insulation

is on the outer surface of the assembly, cold air circulating in this space will “short circuit” the insulation.
It is better to apply a full bed of adhesive or a grid of
beads of adhesive (Fig. 5.3.33). A full adhesive bed is
the preferable method from an adhesion point of view
but where it is on the cold side of the insulation (e.g.,
applying insulation to the interior surface) it may act as
a vapor trap preventing drying of any moisture which
penetrates the interior air/vapor barrier. In this situation therefore the grid approach should be used.
Stick Clips: These are thin metal or plastic pins with
a large perforated flat head at one end. The head is
fastened to the concrete surface with a high quality
adhesive which keys into the perforations. The clips are
applied in a grid pattern, then the insulation is impaled
on the pins and secured in place with a type of spring
washer which is simply pushed over the end of the pin
against the insulation. Sharp “teeth” on the washer
Fig. 5.3.34 Use of stick clips to install rigid insulation.

Rigid Insulation

Fig. 5.3.33 Application of rigid insulation
with adhesives.
Rigid Insulation
Adhesive Daubs

Washers Hold Insultation
to Wall

Not Recommended
Stick Clips Adhered to Wall

Adhesive Grid
Adhesive Backing
Better Method

Washer
Full Adhesive Bed

Recommended Method

448

|

ARCHITECTURAL PRECAST CONCRETE

grip the pin (Fig. 5.3.34). Although this method also
relies on adhesive, the entire surface does not have
to be covered, thus making it easier to clean the sur-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.7 Application of Insulation

Fig. 5.3.35 Furring system.

Where this method is used to apply insulation to the
inside of a wall, the interior finish is applied by screwing or nailing it to the furring members.

Outside Wall

Insulation may be plant or jobsite applied:

Rigid Insulation

1. Mechanical; Most commonly performed at jobsite.
If done in precast concrete plant, see note below.

Mechanical
Fastener
Z-Metal Furring
Drywall Screw
Drywall

face and permitting the use of high performance (and
hence costly) adhesives. As previously discussed, metal
stick clips will reduce the R-value of the insulation. For
example, metal pins representing 0.06% of an insulated panel area can reduce the panel R-value by 6%.
Insulation would retain its full R-value if plastic pins are
used.
Furring Systems: There are a number of types of
plastic, wood, or metal furring which can be applied
over the insulation and fastened, through it, to the
concrete surface. Figure 5.3.35 illustrates one of the
approaches. The furring is usually applied along the
joint between two insulation boards so that one piece
of furring contributes to the support of two insulation
boards. Depending on the size of the insulation boards
and the amount of support required by any subsequent finish, furring may also be applied in the middle
of the insulation boards. Metal furring will decrease
the effectiveness of the insulation and may also require
special preparation of the insulation. The decrease in
the R-value of the insulation due to metal furring can
be determined using Table 5.3.8, page 411.
The insulation may be held in place temporarily prior
to application of the furring by light daubs of adhesive.
These should be very light to avoid holding the insulation away from the surface as discussed above in the
section on adhesives.
The furring can be fastened with powder-driven fasteners or a special type of concrete nail which is driven
into a predrilled hole. The available length of fasteners
usually limits the thickness of insulation to about 4 in.

5

2. Adhesive; As above.
3. Spraying; Normally accomplished at jobsite after
installation. If done in precast concrete plant, see
note below.
4. Poured; Face to be insulated must be face-up during casting. This permits simple application of insulation following concrete casting and initial curing.
Very lightweight concrete should be checked for
variation due to shrinkage to avoid possible delamination. For soft insulations the note below is
also valid.
5. Wet Application; Insulation should have a bondable surface. Shear ties should be used between
concrete and insulation.
Note: For all insulation applied in the precast concrete
plant, by whatever method, the initial cost saving in
application should be weighed against the cost of added protection during handling and transportation and
possible protection against inclement weather. The latter will depend mostly on the type of insulation used.
Good design development of any wall systems takes
a good understanding of all components within that
wall assembly. In cold weather climates, it is desirable
to locate the wall insulation so it mates easily with the
roof insulation. This approach can be achieved with
precast concrete cornice or parapet panels, either single-wythe or in sandwich panels.
Where precast concrete cladding is applied over a
previously erected wall, as would be the case with a
concrete end shear wall (Fig. 5.3.36), it is necessary to
leave holes in such walls for access to the connection
points for the precast concrete panels. Care must be
taken to fill these holes with insulation after the precast concrete panels are installed in order to maintain
the integrity of the envelope’s airtightness and thermal
resistance. The thermal consideration is especially important where the insulation is installed on the outer
surface of the inner wall prior to erection of the precast
concrete panels, when recommended to avoid thermal
bridges at the slabs. One solution is to fill around the
panel connections with pre-packaged, foam-in-place

ARCHITECTURAL PRECAST CONCRETE

|

449

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.7 Application of Insulation / 5.3.8 Precast Concrete Sandwich Panels

Fig. 5.3.36 Cladding erected over previously erected wall.

Firestop

Roof Slab

Furring
Sealant
Furring
Precast Concrete Wall
Rigid Insulation
Adhesive
Drywall
Concrete Wall
Blocked out for Access to Connection
Connection
Sealant

Floor Slab
Sealant
Firestop
Vertical Section

urethane. The effect of these holes on the envelope’s
airtightness will be less of a concern where the approach of treating the interior finish as the primary air
barrier is adopted. This is not to suggest that the holes
should not be properly sealed when this approach is
adopted. They also represent weaknesses in the wall’s
secondary line of defense against rain penetration.

|

Precast Concrete Panel
Sealant
Floor Slab

Exterior Finish

Drip
Vertical Section at Spandrel /Soffit

5.3.8 Precast Concrete Sandwich Panels

energy conservation. In addition to the low thermal
conductivity (high R-value) of the included insulation
layer, concrete sandwich walls include the mass and
heat capacity of concrete, providing thermal damping
(See Section 5.3.5, page 412). A range of R-values can
be obtained by varying the insulation thickness and
material or, in some cases, by varying the unit weight
of the concrete. The effects of thermal damping can
vary with the climate and the building use. However,
a concrete sandwich wall will almost always provide
both reductions and delays of the peak loads affecting
a building. Therefore, even with equal R-values, a concrete sandwich wall will provide greater energy savings
than a wall constructed of lightweight (low mass) materials. This effect is recognized by the major building
codes and can be considered during the panel design.

Precast concrete sandwich wall panels can provide an
aesthetically pleasing, durable exterior finish, a paintready, durable interior surface, and effective thermal
and moisture protection for a building. Such panels normally comprise an exterior layer (or wythe) of
concrete, a layer of rigid, board insulation, an interior
layer of concrete, and a wythe connector system passing through the insulation, tying the layers of concrete
together. If required, the panels can also include an
external air layer so that they can function as part of a
rain-screen system.

In addition to providing insulation for the building,
sandwich panels must resist structural effects, including lateral forces, gravity loads, and temperature effects. Lateral forces may include seismic, wind, soil,
and blast effects. Gravity loads can include self-weight,
as well as loads imposed by floor or roof structures.
Temperature effects arise due to the natural temperature differential that will occur through the thickness
of the sandwich panel, as well as the temperature gradients that must occur through the thickness of each
concrete layer.

Precast concrete sandwich walls are ideally suited for

In general, sandwich panels are considered to be

Access to the back of the panels for sealing the joints
is not a problem where the inner wall is erected after
the precast concrete panels, and where the inner wall
is a steel stud type, or where there is no inner wall. This,
of course, assumes the panel joints are offset from the
slabs and cross-walls or exterior columns.

450

Blocked out for Access to Connection

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.8 Precast Concrete Sandwich Panels

loadbearing or cladding (non-loadbearing) panels.
Both panel types must be designed to resist lateral
forces applied normal to the plane of the panel and
may be designed to resist in-plane forces applied by
roof or floor diaphragms.
Unlike panels with post-installed insulation systems,
precast concrete sandwich panels provide protection
to the insulation layer against flames and heat. They
therefore limit the production of toxic gases during building fires and do not promote the spreading
of flames to adjacent components. Also in contrast
to panels with post-installed insulation systems, precast concrete sandwich panels protect the insulation
against rodent and impact damage. Finally, in contrast

5

to panels with interior, post-installed insulation, properly detailed precast concrete sandwich panels do not
create conditions that support mold growth.
Sandstone-colored precast concrete, with two tones
of aggregate, creates a banding effect on the new 8story county hospital façade, Fig. 5.3.37. Most of the
building (98,000 sq ft) is clad with precast concrete
insulated sandwich wall panels for energy efficiency
and to eliminate condensation. The panels feature a
2 in. exterior wythe, 2 in. of rigid insulation and, a 6
in. interior wythe. The design team wanted a building
that could be maintained without difficulty and that
would retain its consistency of texture and color for
many years.

Fig. 5.3.37
John H. Stroger, Jr. Hospital of Cook County, Chicago, Illinois; Architect CCH Design Group
a joint venture of Loebl Schlossman & Hackl, Inc., McDonough Associates, Inc.,
Globetrotters Engineering Corporation,
and HDR, Inc.;
Photo: C. Scott McDonald
©Hedrich Blessing.
Insulated sandwich wall panels.

ARCHITECTURAL PRECAST CONCRETE

|

451

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.8 Precast Concrete Sandwich Panels

Fig. 5.3.38
Academic Facility for the John F. Kennedy Special Warfare Center, Fort Bragg (Fayetteville), North Carolina;
Architect: LS3P Associates, Ltd., Photo: LS3P Associates, Ltd.

Insulated sandwich wall panels with exposed interior.

Fig. 5.3.39
Morris County Correctional Facility, Morris Township, New Jersey; Architect: Hellmuth Obata & Kassabaum;
Photo: Worth Construction Company, Inc.

Insulated sandwich wall panels with exposed interior.

452

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.8 Precast Concrete Sandwich Panels

The use of insulated architectural precast concrete
panels for a military academic facility, Fig. 5.3.38, afforded several advantages over other types of cladding. It was a cost effective system, providing both the
exterior and interior finish. The interior surface was
given a light broom finish and painted. Also, overall
construction time was reduced due to rapid installation
of the panels.

precast concrete insulated panels incorporated 2 in.
of rigid insulation between a structural wythe of 5 in.
and an architectural wythe of 4 in. Thin desert ironspot
brick was cast in for contrast to the light-beige precast
concrete. Banding which incorporated black glazed
brick was used to mask the horizontal prison cell windows. Cell areas utilized the thermally efficient panel
as the interior surface in inmate cell areas.

Figure 5.3.39 is a six story pre-sentencing facility
for temporary housing of inmates. The building uses
precast concrete insulated concrete sandwich panels
which have an architectural appeal to blend with corporate office buildings in the immediate vicinity. The

Insulated sandwich wall panels faced with thin brick
were substituted for the traditional cavity-wall construction in the school building, Fig. 5.3.40. This allowed the construction time to be reduced by nearly
20% over a brick and block structure. It also avoided

5

Fig. 5.3.40
Jack Britt High School, Fayetteville, North Carolina; Architect: Shuller Ferris Lindstrom & Associates.

Insulated sandwich wall panels painted on the interior.

Fig. 5.3.41 Non-composite and composite insulated panels.
1

3

1

3

1-1
2-2
2

Only recommended above
roof line or below grade

2

2

2

1
1
(a) Non-Composite Panel

Rigid Ties

1-1
Alternate
Arrangement
at Top or
Bottom.

3-3

4

4

4

4

4-4

3-3
Alternate
Arrangement
at Top.
Composite
Action
Achieved
by ties.

3
3
(b) Composite Panel

ARCHITECTURAL PRECAST CONCRETE

|

453

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.8 Precast Concrete Sandwich Panels

the real possibility that a shortage of skilled masons in
the area could seriously stall the project.
Providing the entire façade package in one unit reduces the number of trades and condenses the responsibility into one supplier, providing a cost-effective
solution.
Because of their unique construction, concrete sandwich panels can act as their own insulated foundation
walls, extending directly from the supporting footing.
This is very important in reducing heat losses to the
ground, especially where deep frost lines prevail. Also,
by allowing the roof connection to be contained on
the interior layer only, sandwich panels can provide a
continuous insulation envelope, even at the roof-topanel connection.
Concrete sandwich panels can be designed to act as
fully composite, partially composite, or non-composite
wall elements, Fig. 5.3.41. The amount of composite
action is a function of both the rigidity and the locations of the inter-layer connectors. More rigid connectors allow a greater percentage of the external forces
to be resisted by axial loads within the concrete layers.
As the rigidity of the connector system is reduced, the
portion of external load resisted by axial loads is also
reduced, leading to a reduction in strength. Further,
as the rigidity of the connector system is reduced, the
shear displacement between concrete layer increases,
leading to significantly reduced panel stiffness.
A common rigid connection system comprises distributed steel elements resembling bar joists, with a chord
member embedded in each concrete layer and with
web members crossing the insulation plane. Another
common rigid connector system comprises discrete,
through-thickness solid sections distributed along the
panel length and width.
Although a fully composite panel normally provides the lightest and thinnest wall section for
resistance of lateral or gravity loads, the negative effects of the resulting thermal bridges (reduced thermal performance of panels as well as
creation of potential zones of surface condensation)
must be evaluated by the design team.
Depending on the rigidity of the connector system
(ties or ribs) wythe interaction may be total or partial.
Non-composite panels (Fig. 5.3.41[a]) are those in
which one wythe is supported from the other by rela-

454

|

ARCHITECTURAL PRECAST CONCRETE

tively flexible ties and/or hangers that allow differential
movement of the wythes with changing temperatures
and humidity conditions. Non-composite panels with
an air space allow for ventilation of the outer wythe
and pressure equalization. For non-composite panels,
one wythe is usually assumed to be “structural” and all
loads are carried by that wythe. The structural wythe
is normally thicker and stiffer than the non-structural
wythe, and is usually located on the interior (warm)
side of the panel to reduce thermal stresses due to
temperature variation. Occasionally it may be the exterior wythe, particularly in the case of sculptured panels
such as ribbed panels, that serves the structural function. Note that metal ties or concrete that penetrate
the insulation may reduce the panel’s R-value, however, non-metallic ties are thermally efficient.
For equal overall thickness of panel, a composite element (Fig. 5.3.41[b]) will have greater lateral stiffness.
However, because the deformation of the outer wythe
will affect the inner wythe, experience indicates that
the lateral bowing of composite panels is slightly more
than that of non-composite panels. While the introduction of prestress in both wythes of a composite panel
has no effect on thermal bowing, it can be used to induce an inward bow to counteract the tendency of the
panel to bow outwards, thus improving the behavior
of the element. While this is difficult to calculate, it is
a workable solution used successfully by experienced
precasters.
Panels with full thicknesses of concrete with or without insulation, or openings with surrounding full thicknesses of concrete, are not recommended because:
1. The full thicknesses of concrete act as restraints
between the two concrete layers, each of which
is subjected to significantly differing deformations,
thus developing forces which may lead to cracking
if the panel is not prestressed. This is true of any
composite panel.
2. The full thicknesses of concrete without insulation
act as significant thermal bridges and will reduce the
insulating effectiveness of the panel, as well as possibly causing local condensation and discoloration.
Some precasters have reported successful use of panels with concrete at full thickness at top and bottom
only. Such an arrangement provides less restraint than
a full thickness of concrete on all sides; however, it is
suggested that this be used with caution and based on
previous experience.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.8 Precast Concrete Sandwich Panels

Precasters who advocate the use of non-composite
sandwich panels emphasize the advantage of the structural wythe being protected by the insulation from extremes of temperature, thus minimizing the bowing of
the structural wythe and eliminating thermal stresses
in the structure. The exterior wythe is free to expand
and contract with variations in temperature.

coating (less than ideal) should be used to minimize absorption of water from the fresh concrete, as this can
have an adverse effect on the thermal performance of
the insulation. In all cases, rigid cellular insulation used
in sandwich panels should not be moisture sensitive.
Extruded polystyrene board insulation (XPS) is moisture
resistant.

Precasters who advocate the use of composite panels
point out that the structural wythe of a non-composite
panel must carry all of the applied loads in addition to
the weight of the other wythe thus increasing overall
dimensions, weight and expense. However a composite panel can be designed to share the loads between
the wythes thus reducing dimensions, weight, and
expense.

The physical properties of the insulations typically
used in sandwich panels are listed on Table 5.3.17(a).
The ASTM references of these insulations are listed in
Table 5.3.17(b).

Insulation and concrete thermal properties are previousley discussed in Section 5.3.3. The insulation should
have low water absorption (ideal) or a water-repellent

5

In some facilities, sandwich panels are exposed to
extremely high interior operating temperatures. The
lack of the necessary capability of an insulation to
withstand these temperatures can cause the panel to
fail to perform as intended throughout the lifetime of
the building. For instance, polystyrene insulation has a
relatively low melting temperature. This type of insula-

Table 5.3.17 (a) Properties of Insulation.

Polystyrene

Polyisocyanurate

Expanded

Extruded

Unfaced

Faced

Cellular
Glass

0.7-0.9

1.1-1.4

1.8

1.31.8-2.2 3.0
1.6

2.0-6.0

2.0-6.0

6.7-9.2

Water absorption
(% volume)

<4.0

<3.0

<2.0

<0.3

<3.0

1.0-2.0

<0.5

Comp. strength (psi)

5-10

13-15

25

15-25 40-60 100

16-50

16

65

105 45-140

500

50

Density (pcf)

Tensile strength (psi)

18-25

Linear coeffient
of expansion
(in/in/°F) x 10-6

25-40

Shear strength (psi)

20-35

Flexural strength (psi)

10-25

30-40

25

25-40
—
50

Thermal conductivity
(Btu-in/hr/ft2/ºF)
0.32-0.28 0.26-0.25 0.23
at 75°F
Max. use temp.

50

165 °F

35

50

30-60

1.6-4.6

20-100

50

40-50 60-75 100

50-210

40-50

60

0.20

0.18

0.10-0.15

0.35

165 °F

250 °F

900 °F

Note: 1 lb per cu ft = 16.02 kg/m , 1 psi = 0.006895 MPa; 1 in/in/ºF = 1.800 mm/mm/°C; 1 Btu-in/hr/ft /°F = 0.1442 Wm/m /C, °C = (°F - 32)/1.8.
3

2

2

Table 5.3.17 (b) ASTM Standard References for Various Types of Insulation.

Type of Insulation

ASTM Designation

ASTM Type

Expanded polystyrene

ASTM C-578

Types I, II, VIII, IX, XI

Extruded polystyrene

ASTM C-578

Types IV, V, VI, VII, X

Polyurethane

ASTM C-591

Types 1, 2, 3

Polyisocyanurate

ASTM C-591

Types 1, 2, 3

ARCHITECTURAL PRECAST CONCRETE

|

455

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.8 Precast Concrete Sandwich Panels

tion will begin to shrink and warp when temperatures
reach 165ºF. Selection of a protected polyurethane or
polyisocyanurate insulation with melting temperatures
above 210ºF can prevent possible structural weakness
or thermal instability. Calcium silicate insulations can
withstand higher temperatures. The specifier should
choose the insulation to be compatible with and resistant to the conditions to which it will be exposed.
The required thickness of the insulation will be determined by the thermal characteristics of the material
and the design temperatures of the structure. A minimum thickness of 1 in. is recommended. The deflection
characteristics of the inter-wythe connectors should
be considered in relation to the insulation thickness.
Although one does not necessarily limit the other, the
two must be designed to be compatible.
Wythe connectors should be installed with minimal
voids in the insulation to avoid forming concrete thermal bridges between wythes. Voids should be filled
with insulation. Low conductivity connectors greatly
improve thermal performance.
The maximum thicknesses and sizes of insulation
commercially available, consistent with the shape of
the panel, are recommended. This will minimize joints
in the insulation and the resulting thermal bridges.
Taped (with a tape that is not moisture sensitive) or
glued abutting ends of single layer insulation, or staggered joints with double layer insulation, will minimize
thermal inefficiencies at joints, if the desired thickness
is not available in one sheet.
The insulation itself may be capable of transferring
a certain amount of shear between the wythes, the
value being dependent upon the thickness and properties of the insulation. It may be necessary to provide
measures to break the bond between the insulation
and the concrete wythes of non-composite panels by
physical or chemical methods to eliminate unintended
restraint. This will allow relatively free movement between the wythes for the dissipation of temperature
and other volume change stresses. While such bond
may be destroyed in time, it is strongest at the initial
stages of casting, when the concrete has its least tensile
strength, and therefore most susceptible to cracking.
Panels may be manufactured by incorporating bondbreakers of polyethylene sheeting or reinforced paper
sheets over the insulation, applying form release agents
to the insulation, or by using two layers of insulation
with staggered joints which will allow movement be-

456

|

ARCHITECTURAL PRECAST CONCRETE

tween the two insulation sheets. This intended movement may be inhibited if the layers are not placed in
a level plane. Similarly, the use of sheeting as a bondbreaker can be negated by unevenness in the bottom
layer of concrete and hence the insulation. Under certain conditions, air gaps between the insulation and
the outer wythe also prevent shear transfer.
The use of tape or sheeting to bridge insulation joints
with a single layer of insulation minimizes concrete
bridges between the wythes. Polyethylene sheeting on
the warm side of the insulation also serves as a vapor retarder. In this case, it is necessary to seal around
mechanical ties between the wythes to provide continuity of the vapor retarder. It should be noted that a
3 in. minimum thickness of the inner structural concrete wythe is normally regarded as a satisfactory vapor retarder, provided that it is quality concrete, has
a low water-to-cement ratio and remains crack-free.
See Section 5.3.6 for appropriate placement of vapor
retarders depending on climate.
Wythe minimum thicknesses are dependent upon
structural requirements, finish, reinforcement protection, handling considerations, and past experience.
Wythes should be kept close to equal thicknesses for
composite panels.
In order to minimize the temperature differential
across the thickness of the non-structural wythe (in
non-composite panels), it should be as thin as architectural details will permit. A non-composite panel usually requires a thicker wythe(s) than a composite panel
with the same load and span conditions. The following
limitations are applicable:
1. At the thinnest point, the thickness of the reinforced panels should not be less than 2 in., but
preferably a minimum of 21/2 in. (or 11/2 in. without
reinforcing bar in the area).
2. Thickness should be sufficient to provide proper
reinforcement cover.
3. Thickness should be sufficient to provide required
anchorage of the wythe connector devices.
4. At no point should the thickness be less than three
times the maximum aggregate size.
The thickness of the structural wythe should be determined by structural analysis, and by the need to
accommodate architectural details. In general, the
structural wythe should not be less than 3 in. thick. In
certain cases, a thinner wythe may be successfully used

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.8 Precast Concrete Sandwich Panels

with rather high quantities of reinforcement and with
a higher risk of cracking and bowing. If the wythes are
prestressed, the wythe should not be less than 2 in.
thick. The wythe thickness may be controlled by the
specified fire resistance for the project.
The other limitations listed above for the non-structural wythe also should be considered. Loadbearing
structural wythes are, in most cases, supported at the
bottom edge. They may have a lateral tie near the top
and a mid-height connection to the adjacent panels to
prevent differential bowing. Non-loadbearing composite or non-composite panels can be supported by hanging from suitably designed connections. It is worth noting that top hung panels eccentrically supported will
bow outward less than the bottom supported units.
Panel Size. Sandwich wall panels are available in almost all of the same shapes and sizes as solid panels.
As with solid panels, the larger the panel, the greater
the economy. The maximum size is limited only by the
handling capability of the plant, erection equipment,
transportation restrictions, and the ability of the panel
to resist the applied stresses. Local precasters should
be contacted to verify optimal panel configurations.
Overall sandwich panel thicknesses have varied from
5 in. to greater than 12 in. Insulation thicknesses commonly vary from between 1 and 6 in.
Special procedures which will reduce the differential
shrinkage magnitude, or magnitude of temperature
differential allow for the production of larger panels.
Such procedures include: (1) use of low shrinkage concrete, and (2) jointing of the non-structural wythe. Any
joints should preferably be complete all the way to the
insulation and should be provided at corners of large
openings in the panels.
Wythe Connectors. Wythe connectors serve a variety of functions. If the panel is cast and stripped in
a flat position, the connectors must be capable of resisting the tension created between the wythes during
stripping. The connectors are also used to transfer wind
and seismic forces between the wythes. In composite
panels, the connectors provide resistance to in-plane
bending shear between the wythes. In non-composite
panels, the type and arrangement of connectors are
detailed to minimize in-plane shear transfer so that the
wythes may act independently. Wythe connectors may
also be required to support the weight of the architectural wythe when the wall panel is bearing only on the
structural wythe.

5

Wythe connectors may be used in various combinations. For example, in a composite panel design, solid
blocks of concrete may be used for in‑plane shear
transfer while metal C-ties can be used to prevent the
wythes from separating. Mechanical wythe connectors penetrate the insulation and are bonded to each
wythe.
Fig. 5.3.42 Non-composite and composite insulated panels.

Bent Reinforcing Bars

Carbon Fiber Truss

Expanded Metal

Welded Wire Truss

Shear connectors are used to transfer (in-plane) shear
forces between the two wythes. Because sandwich wall
panels are usually designed as one-way structural elements, shear forces are generated due to longitudinal
bending in the panels. In some cases, the shear connectors may be used to transfer the weight of a non-structural wythe to the structural wythe. Some shear connectors are intentionally stiff in one direction and flexible in
the other. These are called one-way shear connectors.
Examples of these are longitudinal steel wire trusses,
M-ties, flat sleeve anchors, and small diameter bent
bars. Some one-way shear connectors are shown in Fig.
5.3.42.
Tension connectors resist tension only and are not
capable of transferring in-plane shear forces between
the wythes. They are used in non-composite panels to
Fig. 5.3.43 Tension/compression ties.

Z-Tie

Hairpin

Fiber Tie

C-Tie

ARCHITECTURAL PRECAST CONCRETE

|

457

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.3.8 Precast Concrete Sandwich Panels

transfer normal forces between wythes and in composite panels as auxiliary connectors when the spacing of
shear connectors is large. Because these connectors
are unable to transfer shear, their contribution to composite action is usually neglected. Examples of tension
connectors are plastic pins, metal C-ties, hairpins, and
z-ties. These connectors are shown in Fig. 5.3.43.
Wire tie connectors are usually 12 to 14 gauge,
and preferably of stainless steel, Type 304 or 316.
Galvanized metal or plastic ties may also be acceptable. Ties of welded wire fabric and reinforcing bars
are sometimes used. Ties should be arranged or coated so that galvanic reaction between the tie and reinforcement will not occur. In buildings with high relative
humidities, over 60%, it may be desirable to use plastic
ties to avoid condensation at the tie locations. Plastic
ties will maintain the rated R-value of the insulation
and reduce heat flow through the wall. Consideration
may have to be given to the effect of the plastic tie on
the fire resistance of the wall.
General Architectural Design Considerations for
precast concrete sandwich panels are similar to the
design of single wythe architectural precast concrete
panels. However, there are some special considerations
for precast concrete sandwich wall panels.
Bowing in sandwich panels is a deflection caused by
differential wythe shrinkage, eccentric prestress, thermal gradients through the panel thickness, differential
modulus of elasticity between the wythes and creep
from storage of the panels in a deflected position.
These actions cause one wythe to lengthen or shorten
relative to the other. When wythes are interconnected,
such differential wythe movement may result in curvature of the panel, i.e., bowing. Because most sandwich
panels exhibit some degree of composite interaction
(due to shear transfer by either bonded insulation and/
or by the stiffness of wythe connectors), bowing in all
types of sandwich panels is common.
Some useful observations made by those experienced
with composite sandwich panels are:
• Panels bow outwards most of the time.
• Panels heated by the afternoon sun will bow more
than those that are not, i.e., panels on the south
and west elevations will bow more than those on
the east and north elevations.
• Panels bow daily due to transient thermal gradients.
• Sandwich panels experience a greater thermal gradi-

458

|

ARCHITECTURAL PRECAST CONCRETE

ent than solid panels of equal thickness. This is due to
the superior thermal properties of sandwich panels.
• Panels stored in a bowed position will tend to remain in the bowed position after erection. This may
be due to “locked-in” creep.
• Differential shrinkage can occur between the wythes due to relative humidity differences between
interior and exterior exposures.
• Panels containing wythes with different moduli of
elasticity, such as panels with wythes containing
different concrete strengths but with equal levels
of prestress, will bow due to differential shortening
and creep of the wythes after prestress transfer.
In order to maintain integrity of caulking, connections should be detailed so that adjacent panels move
together perpendicular to their plane. The connections
should also be detailed so that volume change forces
do not build up parallel to the plane of the panels.
It is important that the designer realize that any calculation of anticipated sandwich panel bowing is approximate. The exact amount of actual bowing cannot
be determined by calculation. It is essential that all parties understand there will be bowing, that experience
with similarly configured panels is the best method of
predicting the magnitude of bow, and that the panel
connections be detailed accordingly.
For panels with large openings, joints in the outer
wythes at the corners of such opening are desirable.
These joints should preferably be completely through
to the insulation layer and may subsequently be sealed
or treated architecturally, in the same manner as the
joints between panels.
Control joints may be required in large non-composite panels to break the outer wythe into units which
will not craze or crack due to extreme temperature
changes, or shrinkage and creep of the concrete. The
pattern for such control joints becomes an important
architectural feature and aligning such joints with adjacent panels must be done carefully. These can be
minimized by having the real panel joint expressed as a
recess, but this may not be possible if the outer wythe
is already of minimum thickness. Alternatively, the pattern may be varied and only maintained in alternate
panels, so that a small misalignment will not be noticeable. The potential for crazing or cracking and the
need for control joints in the outer wythe can be reduced by prestressing the panels.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.3.8 Precast Concrete Sandwich Panels / 5.4.1 Sustainability Glossary

Fig. 5.3.44 Corner variations of sandwich panels.

90° Butt Joint

45° Mitered Joint

90° Butt Joint Return

5

to suggestions for corner columns. A door or window
frame can be attached to the inside wythe because
most of the bowing movement is confined to the exterior wythe.
Window frames should have thermal breaks between
the exterior frame and the interior frame. Although extruded aluminum window frames are more commonly
used in precast concrete cladding, other framing materials such as aluminum-clad wood, vinyl, fiberglass,
or aluminum with a thermal break, will experience less
heat loss through the frame. A substantial part of the
total heat loss through a window can occur through its
frame. Care must also be taken to avoid placing moisture sensitive materials in contact with concrete.

5.4 SUSTAINABILITY
5.4.1 Glossary
At corners, the bowing of panels perpendicular to
each other may cause unacceptable separation and
possible damage to the joint sealant. It may be desirable to restrain bowing at the corners with one or
more connections between panels or to a corner column. Good corner details are essential and should be
carefully detailed, Fig. 5.3.44. Mitered corners should
have a quirk detail and be restrained with the panels
adequately prestressed or reinforced to resist the restraint forces. The panel-to-panel connections should
be detailed to minimize significant in-plane volume
change restraint forces. Corner panels are not easy to
weatherseal even with returns as the bowing will be
in different planes. In addition, the panel with even a
small return will be stiffer than its neighbor, and both
joints on either side of a corner may suffer. A separate
corner unit, which is not necessarily flush with the adjacent panels, can be effectively used to camouflage
bowing in the two different planes.
If other materials are incorporated in a wall with precast concrete sandwich wall panels, no attempt should
be made to make this interfacing material flush with
the concrete surface, as it is unlikely that this material will act and bow exactly like the concrete panels.
Anything connected or adjacent to the sandwich panels must be able to accommodate bowing movement.
If it is essential that the panels are in the same theoretical plane, it is suggested that they be framed around
with material which is not flush with the walls, similar

Admixture: material, other than water, aggregate,
and hydraulic cement, used as an ingredient of concrete, mortar, grout, or plaster and added to the batch
immediately before or during mixing. Chemical admixtures are most commonly used for freeze-thaw protection, to retard or accelerate the concrete setting time,
or to allow less water to be used in the concrete.
Albedo: solar reflectance; see reflectance.
Building envelope: the components of a building
that perform as a system to separate conditioned space
from unconditioned space.
Calcination: process of heating a source of calcium
carbonate, such as limestone, to high temperatures,
thereby causing a chemical reaction that releases CO2.
This CO2 is not related to the fuel used to heat the
calcium carbonate.
Cement: see portland cement.
Cementitious material (cementing material): any
material having cementing properties or contributing to the formation of hydrated calcium silicate
compounds. When proportioning concrete, the following are considered cementitious materials: portland cement, blended hydraulic cement, fly ash,
ground granulated blast-furnace slag, silica fume,
calcined clay, metakaolin, calcined shale, and rice
husk ash.
Concrete: mixture of binding materials and coarse and
fine aggregates. Portland cement and water are com-

ARCHITECTURAL PRECAST CONCRETE

|

459

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.1 Sustainability Glossary / 5.4.2 Sustainability Concepts

monly used as the binding medium for normal concrete mixtures, but may also contain pozzolans, slag,
and/or chemical admixtures.
Emittance: the ability of the material to emit, or “let
go of” heat.
Green buildings: buildings designed considering the
concepts of sustainable design and reduction of environmental impacts due to site selection, water use, enegy use, materials and resources, the building’s impact
on the environment, and indoor air quality.
Greenhouse gas emissions: emissions that have the
potential to increase air temperatures at the earth’s surface, including carbon dioxide, methane, nitrous oxide,
CFCs, water vapor, and aerosols (particles of 0.001 to
10μm diameter).
Portland cement: Calcium silicate hydraulic cement produced by pulverizing portland-cement clinker, and usually containing calcium sulfate and other
compounds.
Pozzolan: siliceous or siliceous and aluminous materials, like fly ash or silica fume, which in itself possess
little or no cementitious value but which will, in finely
divided form and in the presence of moisture, chemically react in the presence of portland cement to form
compounds possessing cementitious properties.
Reflectance: the ratio of the amount of light or solar
energy reflected from a material surface to the amount
shining on the surface. Solar reflectance includes light
in the visible and ultraviolet range. For artificial lighting,
the reflectance refers to the particular type of lighting
used in the visible spectrum.
Silica fume: very fine noncrystalline silica which is a
byproduct from the production of silicon and ferrosilicon alloys in an electric arc furnace; used as a pozzolan
in concrete.
Slag cement (Ground granulated blast-furnace slag):
a nonmetallic hydraulic cement consisting essentially of
silicates and aluminosilicates of calcium developed in a
molten condition simultaneously with iron in a blast
furnace. Slag cement can be used as a partial replacement or addition to portland cement in concrete.
Supplementary cementitious materials: materials
that when used in conjunction with portland cement
contribute to the properties of hardened concrete
through hydraulic or pozzolanic activity or both.
Sustainability: development that meets the needs of

460

|

ARCHITECTURAL PRECAST CONCRETE

the present without compromising the ability of future
generations to meet their own needs.19 In more tangible terms, sustainability refers to the following: not
compromising future quality of life; remediating environmental damage done in the past; and recognizing
that our economy, environment, and social well-being
are interdependent.
Sustainability rating systems: a set of criteria used
to certify that a construction, usually a building, is sustainable, green, or energy-conserving.
Thermal mass: the storage properties of concrete and
masonry that result in a reduction and shift in peak
energy load for many buildings in many climates, compared to wood or metal frame structures.
Urban heat island: microclimates near urban or suburban areas that are warmer than surrounding areas
due to the replacement of vegetation with buildings
and pavements.

5.4.2 Sustainability Concepts
Sustainability is often defined as development that
meets the needs of the present without compromising the ability of future generations to meet their own
needs. Worldwide, people are currently using 20%
more resources than can be regenerated. In particular,
the U.S. population consumes more resources on a per
capita basis than any other nation.
The environmental impact of constructing and operating buildings in most countries is significant. Consider
that buildings consume 65% of the electricity generated in the U.S. and more than 36% of the primary
energy (such as natural gas); produce 30% of the national output of greenhouse gas emissions; use 12%
of the potable water in the U.S.; and employ 40% of
raw materials (3 billion tons annually) for construction
and operation worldwide.20
Building materials can have a significant effect on the
environmental impact of the construction and operation of a building. Some materials may have to be used
in special configurations, or employ different combinations, to achieve sustainability; the inherent properties
of precast concrete, however, make it a natural choice
for achieving sustainability in buildings. Precast concrete
19 World Commision on Environment and Development, “Report on Our
Common Future,” Oxford University Press, New York, NY, 1987.
20 U.S. Green Building Council, “An Introduction to the U.S. Green Building
Council and the LEED Green Building Rating System,” PowerPoint presentation on the USGBC website, October 2005, www.usgbc.org.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.2 Sustainability Concepts / 5.4.2.2 Cost of building green

contributes to sustainable practices by incorporating integrated design, using materials efficiently, and reducing construction waste, site disturbance, and noise.
Although most consumers are concerned with the
present and future health of the natural environment,
few are willing to pay more for a building, product,
process, or innovation that minimizes environmental
burdens. The concept of sustainability, however, balances sustainable design with cost-effectiveness. Using
integrated design (also called holistic or whole building
approach), a building’s materials, systems, and design
are examined from the perspective of all project team
members and tenants. Energy efficiency, cost, durability
(or service life), space flexibility, environmental impact,
and quality of life are all considered when decisions are
made regarding the selection of a building design.

5.4.2.1 T riple bottom line
The triple bottom line — environment, society, and
economy — emphasizes that economic consequences
are related to environmental and social consequences.
Consequences to society include impacts on employees, communities, and developing countries, as well
as ethics, population growth, and security. Reducing
material, energy, and emissions used by buildings has
impacts far beyond those of the buildings themselves,
such as:
• Using less materials means fewer new quarries are
needed.
• Using less energy means fewer new power plants
need to be constructed, less pollution is emitted
into the air, and dependence on foreign energy
sources is reduced.
• Less emissions to air means a reduction in respiratory conditions, such as asthma.
• Using less water means a reduction in demands on
the infrastructure to find and deliver new sources
of water.
All of these examples indicate how building energy
and utility use affect the local community. These are
especially important since most communities do not
want new power plants, quarries, or landfills built near
them.
The community can also be considered globally.
Carbon dioxide (CO2) emissions in the U.S. were reduced in 2002 for the first time; this reduction, how-

5

ever, was due to a decrease in manufacturing and a
stagnant economy. That same year, China’s production
of CO2 increased by more than the reduction realized
in the U.S., but this increase was primarily due to production of materials consumed by U. S. citizens. Energy
and material consumption, waste, and emissions to air,
land, and water need to be considered from a global as
well as regional perspective in a global market.

5.4.2.2 Cost of building green
A sustainable design can result in reduced project
costs and a building that is energy and resource efficient. Energy and water efficient buildings have lower
operating costs (in the range of $0.60 to $1.50 versus $1.80 per sq ft) and a higher facility value than
conventional buildings.21 Lower energy costs translate into smaller capacity requirements for mechanical
equipment (heating and cooling) and lower first costs
for such equipment. Effective use of daylighting and
passive solar techniques can further reduce lighting,
heating and cooling costs. Reusing materials, such
as demolished concrete for base or fill material, can
reduce costs associated with hauling and disposing of
materials.
When sustainability is an objective at the outset of
the design process, the cost of a sustainable building is
competitive. Often green buildings cost no more than
conventional buildings because of the resource-efficient
strategies used, such as downsizing of more costly mechanical, electrical, and structural systems. Reported increases in first costs for green buildings range from 0 to
2% or more, with costs expected to decrease as project
teams become more experienced with green building
strategies and design.22 Generally, a 2% increase in construction costs will result in a savings of 10 times the
initial investment in operating costs for utilities (energy,
water, and waste) in the first 20 years of the building’s
life.
Buildings with good daylighting and indoor air quality — both common features of sustainable buildings
— have increased labor productivity, worker retention,
and days worked. These benefits contribute directly
to a company’s profits because salaries — which are
about ten times higher than rent, utilities, and maintenance combined — are the largest expense for most
21 U.S. Green Building Council, “An Introduction to the U.S. Green Building
Council and the LEED Green Building Rating System,” PowerPoint presentation on the USGBC website, October 2005, www.usgbc.org.
22 Green Value, Green Buildings Growing Assets, www.rics.org/greenvalue.

ARCHITECTURAL PRECAST CONCRETE

|

461

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.2.2 Cost of building green / 5.4.2.3 Holistic/integrated design

Table 5.4.1 Integration Strategies.

INTEGRATION STRATEGY

SUSTAINABILITY ATTRIBUTE

Use precast concrete panel as
interior surface.

Saves material; no need for additional framing and drywall.

Use hollow-core panels as ducts.

Saves material and energy; eliminates ductwork and charges thermal mass
of panel.

Use thermal mass in
combination with appropriate
insulation levels in walls.

Thermal mass with insulation provides energy benefits that exceed the
benefits of mass or insulation alone in most climates.

Design wall panels to be
disassembled for when building
function changes.

Saves material; extends service life of panels.

Use durable materials.

Materials with a long life cycle and low maintenance will require less
replacement and maintenance during the life of the building.

Use natural resources such
as daylight as a source for
building lighting, trees for
shading, and natural ventilation

Reduces lighting and cooling energy use. Increases indoor air quality and
employee productivity.

Reduce and recycle construction Reduces transportation and disposal costs of wastes. Less virgin materials are
waste.
used if construction waste is recycled for another project.
Use building commissioning
quality control, and inspections
to ensure that building
standards are met.

Energy savings and indoor air quality are most likely attained during the
building life if inspections are made to ensure construction was completed
as designed.

companies occupying office space.23 In schools with
good daylighting and indoor air quality, students have
higher test scores and lower absenteeism.

5.4.2.3 H
 olistic/integrated design
A key tenet of sustainable design is the holistic or
integrated design approach. This approach requires coordinating the architectural, structural, and mechanical
designs early in the schematic design phases to discern
possible system interactions, and then deciding which
beneficial interactions are essential for project success.
For example, a well-insulated building with few windows that face east and west will require less heating
and air-conditioning. This could impact the mechanical design by requiring fewer ducts and registers and
perhaps allow for the elimination of registers along
the building perimeter. Precast concrete walls act as
thermal storage to delay and reduce peak loads, while
also positively affecting the structural design of the
building. Table 5.4.1 provides other integrated design
strategies.
23 U.S. Green Building Council, “Making the Business Case for High
Performance Green Buildings,” www.usgbc.org.

462

|

ARCHITECTURAL PRECAST CONCRETE

A holistic viewpoint will also take into account the
surrounding site environment:
• Are shelters needed for people who take public
transportation to work?
• Can bike paths be incorporated for those who bike
to work?
• Can native landscaping be used to reduce the need
for irrigation?
The eight elements of integrated design are:
1. Emphasize the integrated process.
2. Consider the building as a whole — often interactive, often multi-functional.
3. Focus on the life cycle.
4. Have disciplines work together as a team from the
start.
5. Conduct relevant assessments to help determine
requirements and set goals.
6. Develop tailored solutions that yield multiple benefits while meeting requirements and goals.
7. Evaluate solutions.
8. Ensure requirements and goals are met.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.2.3 Holistic/integrated design / 5.4.3.1 Life cycle cost and service life

Contracts and requests for proposals (RFPs) should
clearly describe sustainability requirements and project
documentation required.24

5.4.2.4 3
 R’s – reduce, reuse, recycle
The 3R’s of reducing waste can be applied to the
building industry.
Reduce the amount of material used and the toxicity of waste materials. Precast concrete can be designed to optimize (or lessen) the amount of concrete
used. Industrial wastes such as fly ash, slag cement,
and silica fume can be used as partial replacements
for cement with certain aesthetic (color) and stripping
time restrictions. Thereby reducing the amount of cement used in concrete. Precast concrete generates a
low amount of waste with a low toxicity. It is generally
assumed that 2% of the concrete at the plant is waste,
but because it is generated at the plant, 95% of the
waste is used beneficially.
Reuse products and containers; repair what can
be reused. Precast concrete panels can be reused
when buildings are expanded. Concrete pieces from
demolished structures can be reused to protect shorelines. Since the precast process is self-contained, formwork and finishing materials are reused. Wood forms
can generally be used 25 to 30 times without major
maintenance while fiberglass, concrete and steel forms
have significantly longer service lives.
Recycle as much as possible, which includes buying products with recycled content. Concrete in
most urban areas is recycled as fill or road base. Wood
and steel forms are recycled when they become worn
or obsolete. Virtually all reinforcing steel is made from
recycled steel. Many cement plants burn waste-derived
fuels such as spent solvents, used oils, and tires in the
manufacture of cement.

5.4.3 L ife Cycle
A life cycle analysis can be done in terms of the economic life cycle cost or environmental life cycle impact. Although the two approaches are different, they
each consider the impacts of the building design over
the life of the building — an essential part of sustainable design. When the energy and resource impacts
of sustainable design are considered over the life of
the building, a sustainable design often becomes more
24 Portland Cement Association, website for sustainable solutions using
concrete, www.concretethinker.com

5

cost-effective. Conversely, when the energy consuming impacts of a low first cost design are considered
over the life of the building, the building may not be
an attractive investment.
Practitioners of sustainable design believe that the
key to sustainable building lies in long-life, adaptable,
low-energy buildings. The durability and longevity of
precast concrete makes it an ideal choice.

5.4.3.1 L ife cycle cost and service life
A life cycle cost analysis is a powerful tool used to
make economic decisions for selection of building materials and systems. This analysis is the practice of accounting for all expenditures incurred over the lifetime
of a particular structure. Costs at any given time are
discounted back to a fixed date, based on assumed
rates of inflation and the time-value of money. A life
cycle cost is in terms of dollars and is equal to the construction cost plus the present value of future utility,
maintenance, and replacement costs over the life of
the building.
Using this widely accepted method, it is possible to
compare the economics of different building alternatives that may have different cash flow factors but
that provide a similar standard of service. The result is
financial information for decision making, which can
be used to balance capital costs and future operation,
repair or maintenance costs. Quite often building designs with the lowest first costs for new construction
will require higher costs during the building life. So,
even with their low first cost, these buildings may have
a higher life cycle cost. Conversely, durable materials,
such as precast concrete, often have a lower life cycle
cost. In the world of selecting the lowest bid, owners
need to be made aware of the benefits of a lower life
cycle cost so that specifications require durable building materials such as precast concrete.
The Building Life-Cycle Cost software from the
National Institute of Standards and Technology (NIST)
provides economic analysis of capital investments, energy, and operating costs of buildings, systems, and
components. The software includes the means to
evaluate costs and benefits of energy conservation
and complies with ASTM standards related to building
economics and Federal Energy Management Program
requirements.
Accepted methods of performing life cycle cost anal-

ARCHITECTURAL PRECAST CONCRETE

|

463

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.3.1 Life cycle cost and service life / 5.4.3.2 Environmental life cycle inventory and life cycle assessment

Fig. 5.4.1 The four phases in the process of developing an LCA.

yses of buildings assume a 20-year life with the building maintaining 80% of its residual value at the end
of this time period. Buildings actually last hundreds of
years if they are not torn down due to obsolescence.
Sustainability practitioners advocate the foundation
and shell of new buildings be designed for a service
life of 200 to 300 years. Allowing extra capacity in the
columns and floors for extra floors and floor loads and
extra capacity in roofs for roof-top gardens adds to the
building’s long term flexibility.
On the other end of the spectrum, real estate speculators plan for a return on investment in 7 years and
generally do not buy into the life cycle cost approach.
Similarly minimum code requirements for energy conserving measures in the building shell are generally for
5 years, meaning initial insulation levels pay for themselves in 5 years. Since it is difficult and costly to add
more insulation to the building shell after it has been
constructed, the 5-year payback for insulation is not
consistent with the life cycle cost associated with 100
year use of buildings.
Advanced building design guidelines from the New
Buildings Institute (www.NewBuildings.org), American
Society for Heating, Refrigerating, and Air-Conditioning
Engineers (ASHRAE) (www.ASHRAE.org), and others
specify insulation levels for those who want to build
cost effective buildings above minimum code levels.
Alternatively, thermal mass and insulation can be included in the life cycle cost analysis to determine costeffective levels. However, this requires whole building
energy analyses to determine annual costs to heat and
cool the building. Economic levels of insulation depend
on the climate, location, and building type.

5.4.3.2 E nvironmental life cycle
inventory and life cycle assessment
A life cycle assessment (LCA) is an environmental assessment of the life cycle of a product. An LCA looks at
all aspects of a product life cycle — from the first stages of harvesting and extracting raw materials from nature, to transforming and processing these raw materi-

464

|

ARCHITECTURAL PRECAST CONCRETE

als into a product to, using the product and ultimately
recycling it or disposing of it back into nature. An LCA
consists of the four phases shown in Fig. 5.4.1.
The LCA of a building is necessary to evaluate the full
environmental impact of a building over its life. Green
buildings rating systems, models such as BEES (www.
bfrl.nist.gov/oae/software/bees.html), and programs
that focus only on recycled content or renewable resources provide only a partial snapshot of the environmental impact a building can leave. An LCA of a building includes environmental effects due to:
• Extraction of materials and fuel used for energy.
• Manufacture of building components.
• Transportation of materials and components.
• Assembly and construction.
• Operation including energy consumption, maintenance, repair, and renovations.
• Demolition, disposal, recycling, and reuse of the
building at the end of its functional or useful life.
A full set of effects includes land use, resource use,
climate change, health effects, acidification, and
toxicity.
An LCA involves a time consuming manipulation
of large quantities of data. A model such as SimaPro
(www.pre.nl/g) provides data for common building
materials and options for selecting LCA impacts. The
Portland Cement Association (PCA) (www.concrete.
org) publishes reports with life cycle inventory (LCI) data
on cement and concrete. All models require a separate
analysis of annual heating, cooling and other occupant
loads using a program such as DOE-2 (http://simulationresearchLBL.gov) or Energy Plus (www.EnergyPlus.
gov).
An LCI is the first stage of an LCA. An LCI accounts
for all the individual environmental flows to and from a
product throughout its life cycle. It consists of the materials and energy needed to make and use a product
and the emissions to air, land, and water associated
with making and using that product.
Several organizations have proposed how an LCA
should be conducted. Organizations such as the
International Organization for Standardization (ISO)
(www.ISO.org), the Society of Environmental Toxicology
and Chemistry (SETAC), (www.SETAC.org), and the
United States Environmental Protection Agency (US
EPA), (www.EPA.gov), have documented standard pro-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.3.2 Environmental life cycle inventory and life cycle assessment / 5.4.3.3 Concrete and concrete products LCI

Fig. 5.4.2 Precast concrete system boundary.

5

LCI that is itself an ingredient to a product. For example, the upstream profile of cement is essentially an LCI
of cement, which can be imported into an LCI of concrete. The LCI of concrete itself can then be imported
into an LCI of a product, such as an office building.
To get the most useful information out of an LCI, precast concrete should be considered in context of its
end-use. For example, in a building, the environmental impact of the building materials is usually dwarfed
by the environmental effects associated with building
operations such as heating, ventilating, cooling, and
lighting.

cedures for conducting an LCA. These procedures are
generally consistent with each other: they are all scientific, transparent, and repeatable.
LCI Boundary. The usefulness of an LCA or LCI
depends on where the boundaries of a product are
drawn. A common approach is to consider all the environmental flows from cradle-to-cradle. For example,
the system boundary in Fig. 5.4.2 shows the most significant processes for precast concrete operations. It
includes most of the inputs and outputs associated
with producing concrete — from extracting raw material to producing mixed concrete ready for placement
in forms. The system boundary also includes the upstream profile of manufacturing cement, as well as
quarrying and processing aggregates, and transporting
cement, fly ash, and aggregates to the concrete plant.
Energy and emissions associated with transporting the
primary materials from their source to the concrete
plant are also included in the boundary. It does not
include, however, upstream profiles of fuel, electricity,
water, or supplementary cementitious materials. This
LCI also does not include form preparation, placing
the concrete in the formwork, curing, and stripping. A
complete precast concrete LCI would include all these
steps.
An upstream profile can be thought of as a separate

The LCI of materials generally do not consider embodied energy and emissions associated with construction of manufacturing plant equipment and buildings,
nor the heating and cooling of such buildings. This is
generally acceptable if their materials, embodied energy and associated emissions account for less than 1%
of those in the process being studied. For example, the
SETAC guidelines indicate that inputs to a process do
not need to be included in an LCI if (i) they are less
than 1% of the total mass of the processed materials
or product, (ii) they do not contribute significantly to a
toxic emission, and (iii) they do not have a significant
associated energy consumption.

5.4.3.3 C oncrete and concrete
products LCI
The data gathered in an LCI is voluminous by nature
and does not lend itself well to concise summaries;
that is the function of the LCA. The data in typical LCI
reports are often grouped into three broad categories:
materials, energy, and emissions. These LCI data do
not include the upstream profiles of supplementary cementitious materials (such as fly ash, silica fume, etc.)
or energy sources (such as fuel and electricity).
Raw Materials. Approximately 1.6 lb (0.73kg) of raw
materials, excluding water, are required to make 1 lb
(0.45kg) of cement.25,26 This is primarily due to the calcination of limestone. In addition to the mixture water,
the LCI assumes that precast concrete consumes 17.5
gallon/yd3 (85 l/m3) of water for washout of the mixer
and equipment used to transfer concrete to molds.
25 Marceau, M.L., Nisbet, M.A., and VanGeem, M.G., “Life Cycle Inventory
of Portland Cement Manufacture,” PCA R&D Serial No. 2095b, Portland
Cement Association, Skokie, Illinois, 2005. www.cement.org
26 Nisbet, M.A., Marceau, M.L., and VanGeem, M.G., “Environmental Life
Cycle Inventory of Portland Cement Concrete,” PCA R&D Serial No. 2095a,
Portland Cement Association, Skokie, Illinois, 2002.

ARCHITECTURAL PRECAST CONCRETE

|

465

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.3.3 Concrete and concrete products LCI

Solid waste from precast concrete plants is insignificant. Waste is about 2.5% of the mass of
concrete used in production. About 95% of this
waste is further beneficially reused through crushing and recycling, resulting in about 0.2 pcf
(3 kg/m3) (about 0.1%) of actual waste.
Fuel and Energy. The amount of energy required to
manufacture or produce a product can be shown in
units of energy, such as joules or Btu’s, or as amounts
of fuel or electricity. Embodied energy per unit volume
of concrete is primarily a function of the cement content of the mixture. For example, cement manufacturing accounts for about 80% of total energy in a 5,000
psi (35MPa) concrete mixture. Energy used in operations at the concrete plant contributes close to 10%,
while aggregate processing and transportation each
contribute about 5%.
The embodied energy of a concrete mixture increases
in direct proportion to its cement content. Therefore,
the embodied energy of concrete is sensitive to the cement content of the mixture and to the assumptions
about LCI energy data in cement manufacturing.
Replacing cement with supplementary cementitious
materials, such as slag cement or silica fume, has the
effect of lowering the embodied energy of the concrete. Fly ash, slag cement, and silica fume do not
contribute to the energy and emissions embodied in
the concrete (except for the small energy contributions
due to slag granulation/grinding, which is included).27
These products are recovered materials from industrial
processes (also called post-industrial recycled materials) and if not used in concrete would use up valuable
landfill space. With a 50% slag cement replacement
for portland cement in a 5,000 psi (35 MPa) mixture,
embodied energy changes from 2.3 to 1.5 GJ/m3 (1.7
to 1.1 MBtu/yd3), a 34% reduction. Fly ash or slag cement replacement of portland cement can also significantly reduce embodied emissions. For instance, a
45% carbon dioxide emissions reduction is achievable
with 50% substitution of slag for portland cement in
a 7,500 psi (50 MPa) precast concrete mixture. Certain
aesthetic (color) and stripping time restrictions apply
when using supplementary cementitious materials.
Embodied energy of reinforcing steel used in concrete is relatively small because it represents only about
27 Marceau, M.L., Gajda, J., and VanGeem, M.G., “Use of Fly Ash in
Concrete: Normal and High Volume Ranges,” PCA R&D Serial No. 2604,
Portland Cement Association, Skokie, Illinos, 2002.

466

|

ARCHITECTURAL PRECAST CONCRETE

1% of the weight in a unit of concrete and it is manufactured mostly from recycled scrap metal. Reinforcing
steel has over 90% recycled content according to the
Concrete Reinforcing Steel Institute (www.crsi.org)
The process for manufacturing reinforcing bars from
recycled steel uses significant energy and should be
considered if the reinforcing bar content is more than
1% of the weight of the concrete.
It is assumed that at a typical site and in a precast
concrete plant, concrete production formwork is reused a number of times through the repetitious nature
of work, so its contribution to an LCI or LCA is negligible. Steel and wood formwork is generally recycled
at the end of its useful life.
When looking at a complicated product, such as an
office building, the categories of fuel and energy are
considered. However, depending on the life span of the
building, the magnitude of energy use due to operations can be quite large. Building energy-use, including
heating, cooling, ventilating, and lighting, is generally
90 to 95% of life cycle energy-use. This means that the
office building life cycle energy is not sensitive to variations in cement manufacturing, concrete production,
or transportation. The embodied energy of the material comprising a building is relatively minor compared
to the building life cycle energy usage. The building
life cycle energy is primarily a function of climate and
building type, not concrete content.
Emissions to Air. The greatest amount of particulate
matter (dust) comes from cement manufacturing and
aggregate production. The single largest contributor
to particulate emissions in both cement manufacturing
and aggregate production is quarry operations (quarry
operations include blasting, haul roads, unloading, and
stockpiling). In cement manufacture, quarry operations
account for approximately 60% of total particulate
emissions. In aggregate production, quarry operations
are responsible for approximately 90% of particulate
emissions. Approximately 30% of the particulate emissions associated with concrete production are from
aggregate production and approximately 60% are embodied in the cement. However, particulate emissions
from quarries are highly variable and sensitive to how
dust is managed on haul roads and in other quarry
operations.
The amounts of carbon dioxide (CO2) and other combustion gases associated with concrete production
are primarily a function of the cement content in the

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.3.3 Concrete and concrete products LCI / 5.4.3.4 Life cycle impact assessment

5

Table 5.4.2 Some Impact Categories for Performing a Life Cycle Assessment.

Bulk waste

Global warming potential

Production capacity of
drinking water

Carcinogens

Hazardous waste

Production capacity of
irrigation water

Climate change

Human toxicity, air

Radiation

Crop growth capacity

Human toxicity, soil

Radioactive waste

Depletion of reserves

Human toxicity, water

Respiratory inorganics

Ecotoxicity soil, chronic Land use

Respiratory organics

Ecotoxicity water,
acute

Life expectancy

Severe morbidity and
suffering

Ecotoxicity water,
chronic

Morbidity

Severe nuisance

Eutrophication

Nuisance

Soil acidification

Fish and meat
production

Ozone depletion

Species extinction

Fossil fuels

Photochemical smog

Wood growth capacity

mixture designs. Emissions of CO2 increase in approximately a one-to-one ratio with the cement content of
concrete. That is, for every additional pound of cement
per cu yd of concrete, there will be an increase in CO2
emissions by approximately 1 lb (0.45kg). Because of
the CO2 emissions from calcination and from fuel combustion in cement manufacture, the cement content of
the concrete mixture accounts for about 90% of the
CO2 emissions associated with concrete production.
Thus, concrete LCI results are significantly influenced
by the cement content of the concrete mixture and the
basis of the CO2 data in the cement LCI.
The fact that cement manufacturing accounts for approximately 70% of fuel consumption per unit volume
of concrete indicates that the amounts of combustion
gases, sulfur dioxide (SO2), and nitrous oxides (NOx),
are sensitive to cement content of the mixture.
Cement kiln dust is a waste product of the cement
manufacturing process and can be used to help maintain soil fertility. An industry-weighted average of 94
lb of cement kiln dust is generated per ton (39 kg per
metric tonne) of cement. Of this about 75 lb (31 kg)
are land-filled and about 19 lb (8 kg) are recycled in
other operations.
Most emissions to air from the life cycle of an office
building come from the use of heating and cooling
equipment, not from the cement or concrete.

5.4.3.4 L ife cycle impact assessment
In the next phase of analysis, the LCI data is assigned
to impact categories and the relative effect of the inventory data within each impact category is weighted.
Among LCA practitioners, this phase is called life cycle
impact assessment, and it consists of category definition, classification, and characterization. Category definition consists of identifying which impact categories
are relevant to the product being studied. Classification
consists of grouping related substances into impact categories. For example, the gases carbon dioxide (CO2),
methane (CH4), and nitrous oxide (N2O) contribute to
climate change; therefore, they can be grouped together in an impact category called climate change. There
are many impact categories to choose from. The categories chosen depend on the goal and scope of the
LCA. Table 5.4.2 lists some possible impact categories.
According to ISO 14041, the only mandatory step in
life cycle impact assessment is characterization. In characterization, weighting factors are assigned according
to a substance’s relative contribution to the impact
category. In terms of global warming potential, one
pound of CH4 is 20 times more potent than one pound
of CO2, and one pound of N2O is 320 times more potent than one pound of CO2. Therefore, in assessing
the potential for global warming, CO2 is assigned a
weighting factor of 1, CH4 a factor of 20, and N2O a
factor of 320. It is important to consider that there is

ARCHITECTURAL PRECAST CONCRETE

|

467

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.3.4 Life cycle impact assessment

no scientific basis for comparing across impact categories. For example, global warming potential cannot be
compared with potential ozone depletion.
The methodology for life cycle impact assessment

is still being developed, and there is no general and
widespread practice at this time or an agreement on
specific methodologies. As a result, it is common to
use several of the available methods to perform the life
cycle impact assessment.

Table 5.4.3 – LEED* Project Checklist: Precast Concrete Potential Points.

POINTS
AVAILABLE

LEED CATEGORY

CREDIT OR PREREQUISITE

Sustainable Sites

Credit 5.1: Site Development, Protect or Restore Habitat

1

Sustainable Sites

Credit 5.2: Site Development, Maximize Open Space

1

Sustainable Sites

Credit 7.1: Heat Island Effect, Non-Roof

1

Energy and Atmosphere

Prerequisite 2: Minimum Energy Performance

—

Energy and Atmosphere

Credit 1: Optimize Energy Performance

Materials and Resources

Credit 1.1: Building Reuse, Maintain 75% of Existing Shell

1

Materials and Resources

Credit 1.2: Building Reuse, Maintain 95% of Existing Shell

1

Materials and Resources

Credit 2.1: Construction Waste Management, divert 50% by weight or
volume

1

Materials and Resources

Credit 2.2: Construction Waste Management, divert 75% by weight or
volume

1

Materials and Resources

Credit 4.1: Recycled Content, the post-consumer recycled content plus
one-half of the pre-consumer content constitutes at least 10% (based
on cost) of the total value of the materials in the project

1

Materials and Resources

Credit 4.2: Recycled Content, the post-consumer recycled content plus
one-half of the pre-consumer content constitutes at least 20% (based
on cost) of the total value of the materials in the project

1

Materials and Resources

Credit 5.1: Local/Regional Materials, Use a minimum of 10% (based on
cost) of the total materials value

1

Materials and Resources

Credit 5.2: Local/Regional Materials, Use a minimum of 20% (based on
cost) of the total materials value

1

Indoor Environmental
Quality

Credit 3.1: Construction Indoor Air Quality Management Plan, During
Construction

1

Innovation and Design
Process

Credit 1.1: Use of high volume supplementary cementitious materials.
Apply for other credits demonstrating exceptional performance

1†

Innovation and Design
Process

Credits 1.2: Apply for other credits demonstrating exceptional
performance

1†

Innovation and Design
Process

Credits 1.3: Apply for other credits demonstrating exceptional
performance

1†

Innovation and Design
Process

Credits 1.4: Apply for other credits demonstrating exceptional
performance

1†

Innovation and Design
Process

Credit 2.l: LEED Accredited Professional

1

1-10

PROJECT TOTALS
*LEED: Leadership in Energy and Environmental Design.
† Up to 4 additional points can be earned, must be submitted and approved (not included in total).
Note: Scoring System: Certified, 26-32 points; Silver, 33-38 points; Gold, 39-51 points; and Platinum, 52-69 points.

468

|

ARCHITECTURAL PRECAST CONCRETE

23

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.4 Green Building Rating Systems / 5.4.4.1 LEED

5.4.4 G
 reen Building Rating Systems
LCI and LCA are valid methods of assessing sustainability, but they are a complex accounting of all materials, energy, emissions, and waste; and their impacts.
Conversely, green building rating systems have gained
popularity because they are comparatively easy to use
and straightforward. Focus groups have shown that
consumers are interested in furthering sustainability
but are unable to define it. Labeling a green building
with LEED, Energy Star or Green Globes certification
sends the message the building is green without having to perform a complex LCI or LCA.

5.4.4.1 L EED
The Leadership in Energy and Environmental Design
(LEED) green building rating system is a voluntary,
consensus-based national standard for developing
high-performance, sustainable buildings. LEED is both
a standard for certification and a design guide for sustainable construction and operation. As a standard, it
is predominantly performance-based, and as a design
guide, it takes a whole-building approach that encourages a collaborative, integrated design and construction process. LEED is administered by the U.S. Green
Building Council (USGBC, www.usgbc.org). LEED-NC28
is a document that applies to new construction and
major renovation projects and is intended for commercial, institutional, and high-rise residential new construction and major renovation.
Essentially, LEED is a point-based system that provides
a framework for assessing building performance meeting sustainability goals. Points are awarded when a
specific intent is met, and a building is LEED certified if
it obtains at least 26 points out of a total availability of
69 points (LEED-NC). The points are grouped into five
categories: (i) sustainable sites, (ii) water efficiency, (iii)
energy and atmosphere, (iv) materials and resources,
and (v) indoor environmental quality. The more points
earned, the “greener” the building. Silver, gold, and
platinum ratings are awarded for at least 33, 39, and
52 points, respectively.
Appropriate use of precast concrete can help a building earn up to 23 points; 26 are required for LEED
certification. Using concrete can help meet minimum
energy requirements, optimize energy performance,
and increase the life of a building. The constituents of
28 “LEED for New Construction,” Version 2.2, United States Green Building
Council, October 2005, www.USGBC.org.

5

concrete can be recycled materials, and concrete itself
can also be recycled. Concrete and its constituents are
usually available locally. These attributes of concrete,
recognized in the LEED rating system, can help lessen a
building’s negative impact on the natural environment.
Points applicable to precast concrete are summarized
in Table 5.4.3 and explained throughout this chapter.
Points must be documented according to LEED procedures to be earned. The USGBC website contains a
downloadable “letter template” that greatly simplifies
the documentation requirements for LEED.
The buildings in the corporate campus for CH2M Hill
in Englewood, CO are framed with a total precast concrete system, including precast concrete shearwalls,
double tees, inverted tee beams and loadbearing exterior walls, Fig. 5.4.3. The buildings are some of the first
total precast concrete office buildings LEED-certified.
The Arizona Departments of Administration and
Environmental Quality (ADOA & ADEQ) project is a
500,000 sq. ft (46,450m2), single contract project
consisting of two architectural precast concrete clad
office buildings and two precast/prestressed concrete
parking structures, Fig. 5.4.4(a and b). The Arizona
Department of Administration (ADOA) is an 185,000
sq. ft (17,187m2), 4-story office building with an 800
space parking structure. The Arizona Department of
Environmental Quality (ADEQ) is a 6-story, 300,000 sq
ft (27,870m2) office building with a 1,000 space parking structure. Both buildings are registered with the
United States Green Building Council’s LEED program.
The 27-story LEED Platinum certified existing office
building in downtown Sacramento, CA, has precast
concrete panels with punched openings, Fig. 5.4.5. The
windows were pre-mounted, glazed, and caulked at
the plant after casting. The precast concrete panels on
the south and west sides of the building have integral
sun shades with a 1 ft (3m) overhang. The building’s
sustainable features can be grouped into three general
categories; air quality; energy conservation and management; and recycling and recycled products.
The project in Fig. 5.4.6 is a USGBC LEED registered
mixed-use development featuring street level retail and
residential condominiums. The structure’s framing consists of 7 in. (175mm) and 12 in. (300mm) loadbearing
walls which support double tees and flat slabs. The precast concrete walls have a combination of sandblasted
and cast-in thin brick finishes. The façade of this one
building has four distinct architectural styles to appear

ARCHITECTURAL PRECAST CONCRETE

|

469

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.4.1 LEED / 5.4.4.2 Energy Star

as four separate and unique buildings. Mechanical,
electrical and plumbing (MEP) accessories, such as conduit boxes, and mechanical and electrical embeds and
openings were cast integrally into the panels.

5.4.4.2 E nergy Star
Energy Star (www.energystar.gov) is a government/
industry partnership designed to help businesses and

consumers protect the environment and save money
through energy efficiency. Energy Star labeling is available for office equipment such as computers and
monitors, appliances such as refrigerators, and residential and commercial buildings. Buildings that meet
certain criteria and achieve a rating of 75 or better in
the Energy Star program are eligible to apply for the
Energy Star (see www.energystar.gov).
The rating consists of a score on a scale of 1 to 100.

All three total precast concrete buildings are LEED certified.

Fig. 5.4.3
CH2M Hill World Headquarters, Englewood, Colorado: Architect: Barber Architecture.

470

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.4.2 Energy Star

The score represents a benchmark energy performance. For example, buildings that score 75 or greater
are among the United States’ top 25%. In addition,
buildings must maintain a healthy and productive indoor environment.
At the present time, five commercial-building types are
eligible for the Energy Star certification: offices, K–12
schools, supermarket/grocery stores, hotel/motels, and
acute care/children’s hospitals. These building types are

5

broken down further into a number of specific occupancies. For example, office buildings include general office,
bank branch, courthouse, and financial center.
Demonstrating conformance is accomplished through
a web-based software tool called Portfolio Manager
(www.energystar.gov). The program hinges on the unbiased opinions of a professional engineer who must
visit the building and verify that data entered about the
building are correct.

Fig. 5.4.4.(a)
Arizona Department of Administration (ADOA), Phoenix, Arizona; Architect: Opus Architects and Engineers;
Photos: Alex Stricker, Stricker LLC.

Fig. 5.4.4.(b)
Arizona Department of Environmental Quality (ADEQ), Phoenix, Arizona;
Architect: Opus Architects and Engineers.

ARCHITECTURAL PRECAST CONCRETE

|

471

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.4.2 Energy Star / 5.4.4.3 Green Globes

Through the Portfolio Manager, the engineer inputs
the building location and energy consumption and
describes its physical and operating characteristics.
Operating characteristics include such things as average weekly occupancy hours, number of occupants,
and number and types of equipment such as personal
computers, refrigeration cases, cooking facilities, and
laundry facilities. Energy consumption is based on all
sources of energy used per month. In addition to energy performance, the engineer is responsible for demonstrating compliance with industry standards on thermal comfort, indoor air quality, and illumination.

First LEED Platinum
certified existing
high-rise.

The professional engineer assessing the building is
expected to give an opinion about the capability of the
building to provide acceptable thermal environmental conditions per ASHRAE Standard 5529 and its capability to supply acceptable outdoor air per ASHRAE
Standard 6230 (see www.ashrae.org). The engineer is
also expected to give an opinion about the capability

Fig. 5.4.5
The Joe Serna Jr. California EPA Headquarters,
Sacramento, California; Architect: A. C. Martin Partners.

Fig. 5.4.6 Bookends, Greenville, South Carolina; Architect:
Johnston Design Group, LLC; Photo: Johnston Design Group, LLC.

of the building to provide minimum illumination levels
per the Iluminance Selection Procedure in the IESNA
Lighting Handbook31 (see www.iesna.org).
29 Amercian Society of Heating, Refrigerating, and Air-Conditioning
Engineers, ASHRAE Standard 55—Thermal Environmental Conditions for
Human Occupancy, Atlanta, GA, www.ASHRAE.org.
30 Amercian Society of Heating, Refrigerating, and Air-Conditioning
Engineers, ASHRAE Standard 62.1-2004—Ventilation for Acceptable
Indoor Air Quality, Atlanta, GA.
31 Illuminating Engineering Society of North America, Illuminating
Engineering Society of North America Lighting Handbook, 9th edition.
December 2000, New York, NY, www.IESNA.org.

472

|

ARCHITECTURAL PRECAST CONCRETE

In addition, Portfolio Manager has the capability to
manage energy data, analyze trends in energy performance (to make budget and management decisions
regarding investments in energy-related projects),
verify building performance, and track the progress of
building improvements.

5.4.4.3 Green Globes
Green Globes is an online, point-based green building rating system administered by the Green Building
Initiative (www.thegbi.org). Many of the points are sim-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.4.3 Green Globes / 5.4.6.1 Fire resistance

ilar to those in LEED, though the point structure differs;
Green Globes has 1000 total points compared with the
69 for LEED-NC. Certification for Green Globes is available at 35% achievement of the total applicable points
compared with LEED at 38% (26 points). It is easier to
obtain certification in Green Globes, however, because
points that are not applicable to the building are subtracted from the total number of applicable points, so
a higher percentage is obtained for those criteria that
are met.

5.4.5 D
 urability
A key factor in building reuse is the durability of the
original structure. Precast concrete panels provide a
long service life due to their durable and low-maintenance concrete surfaces. A precast concrete shell can
be left in place when the building interior is renovated.
Annual maintenance should include inspection and, if
necessary, repair of sealant material.
Modular and sandwich panel construction with concrete exterior and interior walls provide long-term
durability inside and out. Precast concrete construction provides the opportunity to refurbish the building should the building use or function change, rather
than tear it down and start anew. These characteristics
of precast concrete make it sustainable in two ways: it
avoids contributing solid waste to landfills and it reduces the depletion of natural resources and production of
air and water pollution caused by new construction.
LEED Materials Credit 1 in Building Reuse. The
purpose of this credit is to leave the main portion of
the building structure and shell in place when renovating, thereby conserving resources and reducing
wastes and environmental effects of new construction. The building shell includes the exterior skin and
framing but excludes window assemblies, interior
partition walls, floor coverings, and ceiling systems.
This credit should be obtainable when renovating
buildings with a precast concrete façade, because
concrete generally has a long life. This is worth 1
point if 75% of the existing building structure/shell
is left in place and 2 points if 100% is left in place

5.4.5.1 C orrosion resistance
The inherent alkalinity of concrete results in a system
of concrete and reinforcing steel that does not corrode

5

in most environments. The most common reason for
surface spalling of concrete in buildings is corrosion
of reinforcing steel due to inadequate concrete cover.
Precast concrete offers increased resistance to this type
of spalling because reinforcement and concrete are
placed in a plant, with more quality control than cast-inplace construction. This reduces variations in concrete
cover over reinforcing steel and reduces the likelihood
of inadequate cover.

5.4.5.2 Inedible
Vermin and insects cannot destroy concrete because
it is inedible. Some softer construction materials are
inedible but still provide pathways for insects. Due to
its hardness, vermin and insects will not bore through
concrete.

5.4.6 Resistant to Natural Disasters
Concrete is resistant to wind, hurricanes, floods,
and fire. Properly designed precast concrete is resistant to earthquakes and provides blast protection for
occupants.

5.4.6.1 Fire resistance
Precast concrete offers noncombustible construction
that helps contain a fire within boundaries. As a separation wall, precast concrete helps to prevent a fire
from spreading throughout a building or jumping from
building to building. During wild fires, precast concrete
walls help provide protection to human life and the occupant’s possessions. As an exterior wall, concrete that
endures a fire can often be reused when the building
is rebuilt.
The fire endurance of concrete can be determined
based on its thickness and type of aggregate. Procedures
for determining fire endurance of building materials
are prescribed by ASTM E119. Concrete element fire
endurance is generally controlled by heat transmission
long before structural failure, whereas other construction materials fail by heat transmission when collapse
is imminent. So, a 2-hour fire endurance for a precast
concrete wall will most likely mean the wall gets hot
(experiences an average temperature rise of 250 ºF
[140 ºC] or 325ºF [180ºC] at any one point) whereas a
2-hour fire endurance of a frame wall means the wall is
likely near collapse. Concrete helps contain a fire even if
no water supply is available, whereas sprinklers rely on
a problematic water source.

ARCHITECTURAL PRECAST CONCRETE

|

473

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.6.1 Fire resistance / 5.4.7.1 High humidity and wind-driven rain

Details on determining fire resistance of precast concrete walls are provided in Section 5.6.

5.4.6.2 T ornado, hurricane, and wind
resistance
Precast concrete can be economically designed to resistant to tornadoes, hurricanes, and wind. Hurricanes
are prevalent in coastal regions. Tornadoes are particularly prevalent in the path of hurricanes and in the central plains of the U.S.
Case Study: In 1967, a series of deadly tornadoes hit
northern Illinois. Damages at the time were estimated
at $50 million, with 57 people were killed and 484
homes were destroyed. Two precast/prestressed concrete structures, a grocery store and a high school, were
in the direct path of two of the tornadoes, which struck
almost simultaneously. Repairs to the structural system
of the grocery store (limited to a single crack in the
flanges and stem of a beam subjected to uplift) were
less than $200. In the high school, structural damage
was limited to the flange of one double-tee member
(24 ft [7.5 m] of which was broken off by flying debris)
and damaged concrete diaphragm end closures.

5.4.6.3 F lood resistance
Concrete is not damaged by water; concrete that does
not dry out continues to gain strength in the presence
of moisture. Concrete submerged in water absorbs
very small amounts of water even over long periods of
time, and this water does not damage the concrete.
Conversely, building materials such as wood and gypsum wallboard can absorb large quantities of water and
cause moisture related problems. In flood-damaged areas, the concrete buildings are often salvageable.

|

Case study: The 1994 earthquake in Northridge,
California (Richter Scale 6.8), was one of the costliest
natural disasters in U.S. history. Total damage was estimated at $20 billion. Most engineered structures within the affected region performed well, including structures with precast concrete components. In particular,
no damage was observed in precast concrete cladding
due to either inadequacies of those components, or inadequacies of their connections to the building’s structural systems, and no damage was observed in the
precast concrete components used for the first floor or
first-floor support of residential housing. It should be
noted that parking structures with large plan areas—
regardless of structural system—did not perform as
well as other types of buildings.

5.4.7 Weather Resistance
5.4.7.1 High humidity and
wind-driven rain
Precast concrete is resistant to wind-driven rain and
moist, outdoor air in hot and humid climates. Concrete
is impermeable to air infiltration and wind-driven rain.
Moisture that enters a precast concrete building must
come through joints between precast concrete elements. Annual inspection and repair of joints will minimize this potential. More importantly, if moisture does
enter through joints, it will not damage the concrete.

Concrete will only contribute to moisture problems in
buildings if it is enclosed in a building system that does
not let it dry out, trapping moisture between the concrete and other building materials. For instance, impermeable vinyl wall coverings in hot and humid climates
will act as a vapor retarder and moisture can get trapped
between the concrete and wall covering. For this reason, impermeable wall coverings (such as vinyl wallpaper) should not be used in hot and humid climates.

Good practice for all types of wall construction is to
have permeable materials that breathe (are allowed to
dry) on at least one surface and not encapsulate concrete between two impermeable surfaces. Concrete
breathes and will dry out. Therefore, as long as a precast concrete wall is allowed to breathe on at least one
side and is not covered by an impermeable material on
both wall surfaces, the potential for moisture problems
within the wall system is minimal.

5.4.6.4 E arthquake resistance

More information on condensation potential and
moisture control in precast concrete walls is covered
in Section 5.3.

Precast concrete can be designed to be resistant to

474

earthquakes. Earthquakes in Guam, United States
(Richter Scale 8.1); Manila, Philippines (Richter Scale
7.2); and Kobe, Japan (Richter Scale 6.9), have subjected precast concrete buildings to some of nature’s
deadliest forces. Appropriately designed precast concrete framing systems have a proven capacity to withstand these major earthquakes.

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.7.2 Ultraviolet resistance / 5.4.8.3 Albedo (solar relectance)

5.4.7.2 U
 ltraviolet resistance
The ultraviolet (UV) range of solar radiation does not
harm concrete. Using non-fading colored pigments in
concrete retains the color in concrete long after paints
have faded due to the sun’s effects. Precast concrete is
ideal for using pigments because the controlled production allows for replication of color for all panels for
a project (Figs. 5.4.3 and 5.4.4).

5.4.8 M
 itigating the Urban Heat
Island Effect
Precast concrete provides reflective surfaces that minimize the urban heat island effect. Cities and urban
areas are 3 to 8 °F (2 to 4 °C) warmer than surrounding areas due to the urban heat island effect. This difference is attributed to heat absorption of buildings
and pavements that have taken the place of vegetation. Trees provide shade that reduces temperatures at
the surface. Trees and plants provide transpiration and
evaporation that cool the surfaces and air surrounding
them. Research has shown the average temperature of
Los Angeles has risen steadily over the past half century, and is now 6 to 7 °F (3 to 4 °C) warmer than 50
years ago.32

5.4.8.1 W
 armer surface temperatures
Urban heat islands are primarily attributed to horizontal surfaces, such as roofs and pavements, that absorb solar radiation. In this context, pavements include
roads, streets, parking lots, driveways, and walkways.
Vertical surfaces, such as the sides of buildings, also
contribute to this effect. Using materials with higher
albedos, such as concrete, will reduce the heat island
effect, save energy by reducing the demand for air
conditioning, and improve air quality (Fig. 5.4.7).
Studies indicate people will avoid using air-conditioning at night if temperatures are less than 75 °F (24 °C).
Mitigating the urban heat island effect to keep summer temperatures in cities less than that temperature
at night has the potential to save large amounts of
energy by avoiding air-conditioning use.

5.4.8.2 S mog
Smog levels have also been correlated to temperature
rise. Thus, as the temperature of urban areas increases,
32 Heat Island Group Home Page, eetd.lbl.gov/HeatIsland/.

5

so does the probability of smog and pollution. In Los
Angeles, the probability of smog increases by 3% with
every degree Fahrenheit of temperature rise. Studies
for Los Angeles and 13 cities in Texas have found that
there are almost never any smog episodes when the
temperature is below 70 °F (21 °C). The probability of
episodes begins at about 73 °F (23 °C) and, for Los
Angeles, exceeds 50% by 90 °F (32 °C). Reducing the
daily high in Los Angeles by 7 °F (4 °C) is estimated to
eliminate two-thirds of the smog episodes.
Smog and air pollution are the main reasons EPA
mandates expensive, clean fuels for vehicles and reduced particulate emissions from industrial facilities
such as cement and asphalt production plants. The
EPA now recognizes that air temperature is as much a
contributor to smog as nitrogen oxide (NOx) and volatile organic compounds (VOCs). The effort to reduce
particulates in the industrial sector alone costs billions
of dollars per year, whereas reduction in smog may
be directly related to the reflectance and colors of the
infrastructure that surround us. Installing low-albedo
roofs, walls, and pavements is a cost-effective way to
reduce smog.

5.4.8.3 Albedo (solar reflectance)
Albedo, which in this case is synonymous with solar
reflectance, is the ratio of the amount of solar radiation reflected from a material surface to the amount
shining on the surface. Solar radiation includes the
ultraviolet as well as the visible spectrum. Albedo is
measured on a scale not reflective (0.0) to 100% reflective (1.0). Generally, materials that appear to be
light-colored in the visible spectrum have high albedo
and those that appear dark-colored have low albedo.
Because reflectivity in the solar radiation spectrum determines albedo, color in the visible spectrum is not
always a true indicator of albedo.
Surfaces with lower albedos absorb more solar radiation. The ability to reflect infrared light is of great importance because infrared light is most responsible for
heating. On a sunny day when the air temperature is
55 °F (13 °C), surfaces with dark acrylic paint will heat
up to 90 °F (32 °C) more than air temperatures, to 145
°F (63 °C). Light surfaces, such as white acrylic, will
heat up to 20 °F (11 °C) more, to a temperature of 75
°F (24 °C). The color and composition of the materials
greatly affect the surface temperature and the amount
of absorbed solar radiation. The effect of albedo and

ARCHITECTURAL PRECAST CONCRETE

|

475

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.8.3 Albedo (solar relectance)

Fig. 5.4.7
Cape Coral City Hall; Cape Coral Florida;
Architect: Spillis Candela/DMJM; Photo: Spillis Candela/DMJM.

High-reflecting (usually light-colored) surfaces help mitigate urban heat islands.

solar radiation on surface temperatures is referred to
as the sol-air temperature and can be calculated.
Traditional portland cement concrete generally has an
albedo or solar reflectance of approximately 0.4, although values can vary; measured values are reported
in the range of 0.4 to 0.5. The solar reflectance of new
concrete is greater when the surface reflectance of the
sand and cementitious materials in the concrete are
greater. Surface finishing techniques also have an effect, with smoother surfaces generally having a higher

476

|

ARCHITECTURAL PRECAST CONCRETE

albedo. For concrete elements with “white” portland
cement, values are reported in the range of 0.7 to 0.8.
Albedo is most commonly measured using a solar-spectrum reflectometer (ASTM C 1549)33 or a pyranometer
(ASTM E 1918).34
33 American Society for Testing and Materials, ASTM C 1549, “Standard
Test Method for Determination of Solar Reflectance Near Ambient
Temperature Using a Portable Solar Reflectometer,” Conshohocken, PA,
www. ASTM.org.
34 American Society for Testing and Materials, ASTM E 1918, “Standard Test
Method for Measuring Solar Reflectance of Horizontal and Low-Sloped
Surfaces in the Field,” West Conshohocken, PA.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.8.4 Emittance / 5.4.9.1 Radiation and toxicity

5.4.8.4 E mittance

5.4.8.6 Mitigation approaches

In addition to albedo, the material’s surface emittance
affects surface temperature. While albedo is a measure
of the solar radiation reflected away from the surface,
surface emittance is the ability of the material to emit,
or “let go of” heat. A white surface exposed to the
sun is relatively cool because it has a high reflectivity
and a high emittance. A shiny metal surface is relatively
warm because it has a low emittance, even though it
has a high albedo. The emittance of most non-reflecting (non-metal) building surfaces such as concrete is in
the range of 0.85 to 0.95. The emittance of aluminum
foil, aluminum sheet, and galvanized steel, all dry and
bright, are 0.05, 0.12, and 0.25, respectively.

One method to reduce the urban heat island effect is
to change the albedo of the urban area. This is accomplished by replacing low albedo surfaces with materials
of higher albedo. This change is most cost effective
when done in the initial design or during renovation
or replacement due to other needs. Planting trees for
shade near buildings also helps mitigate the urban
heat island effect. Shade also directly reduces the airconditioning load on buildings. Using deciduous trees
shades the buildings in the summer and allows the sun
to reach the buildings in the winter.

5.4.8.5 M
 oisture
Moisture in concrete helps to cool the surface by
evaporation. Concrete when placed has a moisture
content of 100% relative humidity. The concrete surface gradually dries over a period of one to two years
to reach equilibrium with its surroundings. Concrete
surfaces exposed to rain and snow will continue to be
wetted and dried. This moisture in the concrete surface
will help to cool the concrete by evaporation whenever
the vapor pressure of the moisture in the surface is
greater than that of the air. In simpler terms, when the
temperature and relative humidity of the air are greater
than that just beneath the concrete surface, the concrete will dry and cool somewhat by evaporation.
The albedo of concrete decreases when the surface
is wet. Consequently, albedo is lower when concrete
is relatively new and the surface has not yet dried,
and when the concrete becomes wet. The albedo of
new concrete generally stabilizes within two to three
months.
LEED Sustainable Sites Credit 7.1 on Heat Island
Effect, Non-Roof. The intent of this credit is to
reduce heat islands. The requirements are met by
placing a minimum of 50% of parking places underground or covered by a parking structure. Precast
concrete parking structures, can be used to help obtain this point. Any roof used to shade or cover parking must meet specified criteria. This credit is worth
1 point.

5

5.4.8.7 Thermal mass and nocturnal
effects
The thermal mass of concrete delays the time it takes
for a surface to heat up but also delays the time to cool
off. For example, a white non-concrete roof will get
warm faster than concrete during the day, but will also
cool off faster at night. Concrete surfaces are often
warmer than air temperatures in the evening hours.
Concrete’s albedo and thermal mass will help mitigate
heat island effects during the day but may contribute
to the nocturnal heat island effect. The moisture absorbed by concrete during rain events helps reduce
the daytime and nocturnal heat island effect when it
evaporates. The challenge is to use concrete to mitigate heat islands while keeping evening temperatures
as cool as possible.

5.4.9 Environmental Protection
5.4.9.1 Radiation and toxicity
One goal of sustainability is to reduce radiation and
toxic materials; concrete provides an effective barrier against radiation and can be used to isolate toxic
chemicals and waste materials. Concrete protects
against the harmful effects of X-rays, gamma rays, and
neutron radiation.
Concrete is resistant to most natural environments;
it is sometimes exposed to substances that can attack
and cause deterioration. Concrete in some chemical
manufacturing and storage facilities must be specifically designed to avoid chemical attack. The resistance
of concrete to chlorides is good, and using less permeable concrete can increase the resistance even more.
This is achieved by using a low water-to-cementitious

ARCHITECTURAL PRECAST CONCRETE

|

477

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.9.1 Radiation and toxicity / 5.4.10 Precast Concrete Production

materials ratio (around 0.40), adequate curing, and
supplementary cementitious materials such as slag cement or silica fume. The best defenses against sulfate
attack where this is an issue, are the measures suggested previously, in addition to using cement specially
formulated for sulfate environments.

5.4.9.2 R
 esistance to noise
Precast concrete walls provide a buffer between outdoor noise and the indoor environment. Because land is
becoming scarcer, buildings are being constructed closer together and near noise sources such as highways,
railways, and airports. The greater mass of concrete
walls can reduce sound penetrating through a wall by
over 80% compared to wood or steel frame construction (Section 5.5). Although some sound will penetrate
the windows, a concrete building is often two-thirds
quieter than a wood or steel frame building.
Precast concrete panels also provide effective sound
barriers separating buildings from highways or industrial areas from residential areas.

5.4.9.3 S ecurity and impact resistance
Concrete is often used as a first line of defense
against explosions or blasts. Rows of concrete planters
or bollards are now positioned in front of most federal
buildings such as court houses, office buildings. and
other high-security areas. Decorative concrete walls
are also used as a primary line of defense to prevent
vehicles from coming close to buildings. From a holistic
perspective, the barriers may also provide benches and
a visual separation between the street and plaza.

5.4.10 P
 recast Concrete Production
The production of precast concrete has many environmental benefits, including:
1. Less materials are required because precise mixture
proportions and tighter tolerances are achievable.
2. Optimal insulation levels can be incorporated into
precast concrete sandwich panel walls.
3. Less concrete waste is created because of tight
control of quantities of constituent materials.
4. Waste materials are more likely to be recycled
because concrete production is in one location.
a. Gray water often recycled into future mixtures.
b. Hardened concrete recycled (presently about 5

478

|

ARCHITECTURAL PRECAST CONCRETE

to 20% of aggregate in precast concrete can
be recycled concrete; in the future this could
be higher.)
c. Steel forms and other materials are reused.
5. Less dust and waste is created at the construction
site because only needed precast concrete elements are delivered and there is no debris from
formwork and associated fasteners—construction
sites are cleaner and neater.
6. Fewer trucks and less time are required for construction because concrete is made offsite; this is
particularly beneficial in urban areas where minimal traffic disruption is critical.
7. Precast concrete units are normally large components, so greater portions of the building are completed with each activity.
8. Less noise at construction site because concrete is
made offsite.
made offsite.
LEED Sustainable Sites Credit 5.1 on Site
Development, Protect, or Restore Habitat. The
intent of this credit is to encourage the conservation
of natural areas on the site and restore damaged
areas. The requirements are met by limiting site disturbance to prescribed distances. Tuck-under parking, such as precast concrete parking structures, can
be used to help obtain this credit worth 1 point.
Also precast concrete requires minimal site disturbance for erection of panels.

LEED Sustainable Sites Credit 5.2 on Site
Development, Maximize Open Space. The intent of this credit is to provide a high ratio of open
space to development footprint. The requirements
are met by limiting the size of the development
footprint; specifically, by exceeding the local zoning’s open space requirement for the site by 25%.
Tuck-under parking, such as precast concrete parking structures, can be used to help obtain this credit
worth 1 point.

Less concrete is generally used in precast concrete
buildings than in other concrete buildings because
of the optimization of materials. A properly designed
precast concrete system will result in smaller structural members, longer spans, and less material used
on-site; this translates directly into economic savings,

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.10 Precast Concrete Production / 5.4.10.1 Constituent materials

which can also result in environmental savings. Using
less materials means using fewer natural resources and
less manufacturing and transportation energy—not to
mention the avoided emissions from mining, processing, and transporting raw and finished material.
Concrete products can provide both the building structure, and the interior and exterior finishes. Structurally
efficient columns, beams, and slabs can be left exposed
with natural finishes. Interior and exterior concrete
walls offer a wide range of profile, texture, and color
options that require little or no additional treatment to
achieve aesthetically pleasing results. Exposed ceiling
slabs and architectural precast concrete panels are some
examples of this environmentally efficient approach.
This structure/finish combination reduces the need for
the production, installation, maintenance, repair, and
replacement of additional finish materials. It also eliminates products that could otherwise degrade indoor air
quality. This approach provides durable finishes that are
not prone to damage or fire. Exposing the mass of the
structure moderates heating and cooling loads.

5.4.10.1 C onstituent materials
Concrete. Concrete is basically a mixture of two components: aggregates and paste. The paste, comprised
of portland cement and water, binds the aggregates
(usually sand and gravel or crushed stone) into a rocklike mass. The paste hardens because of the chemical
reaction of the cement and water. Supplementary cementitious materials and chemical admixtures may also
be included in the paste. The absolute volume of cement is usually between 7% and 15% and the water
between 14% and 21%.
Portland Cement. Portland cement (hereafter called
cement) is made by heating common minerals, primarily crushed limestone, clay, iron ore, and sand, to
a white-hot mixture to form clinker. This intermediate
product is ground, with a small amount of gypsum,
to form a fine gray powder called cement. To trigger
the necessary chemical reactions in the kiln, these raw
materials must reach about 2700°F (1482ºC)–the temperature of molten iron. Although the portland cement
industry is energy intensive, the U.S. cement industry
has reduced energy usage per ton of cement by 35%
since 1972.35,36
35 Portland Cement Association, U.S. and Canadian Labor-Energy Input
Survey, Skokie, IL,www.cement.org.
36 Portland Cement Association, “Report on Sustainable Manufacturing”,
2006,www.cement.org.

5

Carbon dioxide emissions from a cement plant are
divided into two source categories: combustion and
calcination. Combustion accounts for approximately
35% and calcination 65% of the total CO2 emissions
from a cement manufacturing facility. The combustion-generated CO2 emissions are related to fuel use.
The calcination CO2 emissions are formed when the
raw material is heated and CO2 is liberated from the
calcium carbonate. As concrete is exposed to the air
and carbonates, it reabsorbs some of the CO2 released
during calcination. Calcination is a necessary key to cement production. Therefore, the focus of reductions
in CO2 emissions during cement manufacturing is on
reducing fuel and energy use.
White portland cement is a true portland cement that
differs from gray cement chiefly in color. The manufacturing process is controlled so that the finished product
will be white. White portland cement is made of selected raw materials containing negligible amounts of
iron and magnesium oxides–the substances that give
cement its gray color. White cement is used primarily
for architectural purposes in structural walls, precast
concrete, and glass fiber reinforced concrete (GFRC)
facing panels. Its use is recommended wherever white
or colored concrete, grout, or mortar is desired. White
portland cement should be specified as white portland
cement meeting the specifications of ASTM C 150,
Type I, II, III, or V.
Abundant Materials. Concrete is used in almost
every country of the world as a basic building material. Aggregates, about 85% of concrete, are generally
low-energy, local, naturally occurring sand and stone.
Limestone and clay needed to manufacture cement are
prevalent in most countries. Concrete contributes to a
sustainable environment because it does not use scarce
resources. Limestone and aggregate quarries are easily reused. While quarrying is intense, it is closely contained and temporary. When closed, aggregate quarries are generally converted to their natural state or
into recreational areas or agricultural uses. In contrast,
other material mining operations can be extensive and
involve deep pits that are rarely restored, and deforestation can have negative environmental effects.
Fly Ash, Slag Cement, and Silica Fume. Fly ash,
slag cement, and silica fume are industrial by-products;
their use as a replacement for portland cement does
not contribute to the energy and CO2 effects of cement in concrete. If not used in concrete, these sup-

ARCHITECTURAL PRECAST CONCRETE

|

479

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.10.1 Constituent materials

plementary cementitious materials (SCMs) would use
valuable landfill space. Fly ash (Fig. 5.4.8 [a]) is a byproduct of the combustion of pulverized coal in electric
power generating plants. Slag cement (Fig. 5.4.8 [b]) is
made from iron blast-furnace slag.37 Silica fume (Fig.
(a)

fly ash

5.4.8 [c]) is a by-product from the electric arc furnace
used in the production of silicon or ferrosilicon alloy.
These types of industrial by-products are considered
post-industrial or pre-consumer recycled materials. Fly
ash is commonly used at cement replacement levels up
to 25%, slag cement up to 60%, and silica fume up to
5 to 7%. When slag cement replaces 50% of the portland cement in a 7500 psi (50 MPa) concrete mixture,
greenhouse gas emissions per cu. yd. of concrete are
reduced by 45%. Because the cementitious content of
concrete is about 15%, these pozzolans typically account for only 2 to 5% of the overall concrete material
in buildings.
SCMs may slightly alter the color of hardened concrete. Color effects are related to the color and amount
of the material used in concrete. Many SCMs resemble
the color of portland cement and therefore have little
affect on color of the hardened concrete. Some silica
fumes may give concrete a slightly bluish or dark gray
tint and tan fly ash may impart a tan color to concrete when used in large quantities. Slag cement and
metakaolin (a clay SCM without recycled content) can
make concrete lighter. Slag cement can initially impart
a bluish or greenish undertone that disappears over
time as concrete is allowed to dry.

(b)

slag cement

silica fume

(c)

white
silica fume

Fig. 5.4.8(a), (b) & (c)
37 Slag Cement Association, “Slag Cement and the Environment,” Slag
Cement in Concrete No. 22, 2003, www. slagcement.org.

480

|

ARCHITECTURAL PRECAST CONCRETE

The optimum amounts of supplementary cementing
materials used with portland or blended cement are
determined by testing, the relative cost and availability
of the materials, and the specified properties of the
concrete. When supplementary cementitious materials are used, the proportioned concrete mixture (using
the project materials) should be tested to demonstrate
that it meets the required concrete properties for the
project. Some pozzolans increase curing times, but this
is not as great a concern for precast concrete manufacturing as it is in cast-in-place applications where construction schedule has a greater impact.
The durability of products with recycled content materials should be carefully researched during the design
process to ensure comparable life cycle performance.
There would obviously be a net negative impact if a
product offering a 20 to 30% recycled content had
only half the expected service life of a product with a
lower or no recycled content.
Recycled Aggregates. The environmental attributes
of concrete can be improved by using aggregates
derived from industrial waste or using recycled con-

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.10.1 Constituent materials

crete as aggregates. Blast furnace slag is a lightweight
aggregate with a long history of use in the concrete
industry.
Recycled concrete can be used as aggregate in new
concrete, particularly the coarse portion. When using the recycled concrete as aggregate, the following
should be taken into consideration:
1. R
 ecycled concrete as aggregate will typically have
higher absorption and lower specific gravity than
natural aggregate and will produce concrete
with slightly higher drying shrinkage and creep.
These differences become greater with increasing
amounts of recycled fine aggregates.
2. T oo many recycled fines can also produce a harsh
and unworkable mixture. Many transportation
departments have found that using 100% coarse
recycled aggregate, but only about 10 to 20% recycled fines, works well.38 The remaining percentage of fines is natural sand.
3. W
 hen crushing the concrete (Fig. 5.4.9), it is difficult to control particle size distribution, meaning
that the “aggregate” may fail to meet grading requirements of ASTM C33.39
4. The chloride content of recycled aggregates is of
concern if the material will be used in reinforced
concrete. This is particularly an issue if the recycled
concrete is from pavements in northern climates
where road salt is freely spread in the winter.
5. The alkali content and type of aggregate in the system is probably unknown, and therefore if mixed
with unsuitable materials, a risk of alkali-silica reaction (ASR) is possible.

Fig. 5.4.9 Crushed concrete from other sources can serve as
recycled aggregate. Photo: Portland Cement Association.
38 Portland Cement Association, Design and Control of Concrete Mixes,
Chapter 5, EB001, 2002, Skokie, IL.
39 American Society for Testing and Materials, ASTM C 33, “Standard
Specification for Concrete Aggregates,” West Conshohocken, PA,
www.ASTM.org.

5

LEED Materials Credit 4 on Recycled Content.
The requirements of this credit state: “Use materials
with recycled content such that post-consumer recycled content plus one-half of the pre-consumer content constitutes at least 10% (based on cost) of the
total value of the materials in the project.” The percentage is determined by multiplying the price of an
item by the percent of recycled materials—on a mass
basis—that make up that item. To earn this credit,
the project must meet the threshold percentages
based on the total of all permanently installed building materials used on the project. Supplementary cementitious materials, such as fly ash, silica fume, and
slag cement, are considered pre-consumer. Since
the cementitious content of concrete is about 15%,
these pozzolans typically account for only 2 to 5%
of the overall concrete material in buildings. For this
reason, LEED-NC v2.2 allows the recycled content of
concrete to be based on the recycled content of the
cementitious materials. Using recycled concrete or
slag as aggregate instead of extracted aggregates
qualifies as post-consumer. Although most reinforcing bars are manufactured from recycled steel, in
LEED, reinforcement is not considered part of concrete. Reinforcing material should be considered as
a separate item. This credit is worth 1 point for the
quantities quoted above and 2 points for double the
amount.
LEED Innovation Credit on Reducing Cement
Content. LEED has an innovation credit that allows
1 point for a 40% reduction of cement content
compared to common practice. This can be met by
using at least 40% less portland cement or replacing at least 40% of the cement in concrete with
fly ash, slag cement, silica fume, or a combination
of the three. Slag cement is commonly used at replacement levels up to 60%. However, fly ash replacement levels for portland cement greater than
25% are not common, as the fly ash and portland
cement need to be chemically and physically compatible to ensure durable quality concrete that sets
properly. For quality concrete, mixtures with fly ash
at replacement levels greater than 25% should not
be used without proven field experience or laboratory testing. Certain aesthetic (color) and stripping
time restrictions will apply when using supplementary cementitious materials.

ARCHITECTURAL PRECAST CONCRETE

|

481

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.10.1 Constituent materials / 5.4.11.1 Energy codes

Admixtures. The freshly mixed (plastic) and hardened
properties of concrete may be changed by adding
chemical admixtures to the concrete, usually in liquid form, during batching. Chemical admixtures are
commonly used to (1) adjust setting time or hardening, (2) reduce water demand, (3) increase workability,
(4) intentionally entrain air, and (5) adjust other fresh
or hardened concrete properties. Admixtures provide
enhancing qualities in concrete but are used in such
small quantities that they do not adversely affect the
environment. Their dosages are usually in the range of
0.005 to 0.2% of the concrete mass.
Color Pigments. Non-fading color pigments are used
to provide the decorative colors in precast concrete. They
are insoluble and generally non-toxic, although some
may contain trace amounts of heavy metals. Many iron
oxide pigments are primarily the byproduct of material
recycling (manufactured by precipitating scrap steel).
Local Materials. Using local materials reduces the
transportation required to ship heavy building materials,
and the associated energy and emissions. Most precast
concrete plants are within 200 miles (300 km) of a building site. The cement, aggregates, and reinforcing steel
used to make the concrete and the raw materials to
manufacture cement are usually obtained or extracted
from sources within 200 miles of the precast concrete
plant. The primary raw materials used to make cement
and concrete are abundant in all areas of the world.
Precast concrete elements are usually shipped efficiently because of their large, often repetitive sizes and
the ability to plan their shipment during the normal
course of the project.

5.4.11 Energy Use in Buildings
Energy conservation is a key tenet of sustainability.
About 90% of the energy used during a building’s life
is attributed to heating, cooling, and other utilities. The
remaining 10% is attributed to manufacturing materials, construction, maintenance, replacement of components, and demolition.40 Approximately 5% of the
world’s population resides in the U.S., yet 25% of the
world’s energy is consumed in the U.S. The U.S. dependence on foreign energy sources is greater than ever,
which has an effect on U.S political and defense policies. Meanwhile, many developing nations like China
have increased energy demands due to increased manufacturing and urbanization.

5.4.11.1 Energy codes
Energy codes provide cost effective, minimum building requirements that save energy. The energy saved
is a cost savings through lower monthly utility bills,
and smaller, and thus less expensive HVAC equipment.
More than two-thirds of the electricity and one-third of
the total energy in the U.S. are used to heat, cool, and
operate buildings.41 This means that implementing and
enforcing energy codes will result in fewer power plants
and natural resources being used to provide electricity and natural gas. It also means fewer emissions will
be released into the atmosphere. Emissions have been
40 Marceau, M.L. and VanGeem, M.G., “Modeling Energy Performance of
Concrete Buildings for LEED-NCv2.1 EA Credit 1,” PCA R&D Serial No.
2880a, Portland Cement Association, Skokie, Illinois, 2006, www.cement.
org.
41 “An Introduction to the U.S. Green Building Council and the LEED Green
Building Rating System,” a PowerPoint presentation on the USGBC website, October 2005, www.usgbc.org.

LEED Materials Credit 5 on Regional Materials. The requirements of this credit state: “Use building materials
or products that have been extracted, harvested, or recovered, as well as manufactured, within 500 miles (800
km) of the project site for a minimum of 10% (based on cost) of the total materials value.” This means that a
precast concrete plant within 500 miles of the building would qualify if the materials to make the concrete were
extracted within 500 miles. Calculations can also include concrete either manufactured or extracted locally.
Precast concrete will usually qualify because precast concrete plants are generally within 200 to 500 miles (300 to
800 km) of a project. Precast concrete plants generally use aggregates that are extracted within 50 miles (80 km) of
the plant and within 200 to 500 miles of the project. Cement and supplementary cementitious materials used for
buildings are also primarily manufactured within 500 miles of a project. Reinforcing steel is also usually manufactured within 500 miles of a project and is typically made from recycled materials from the same region.
Using materials that are extracted or manufactured locally supports the regional economy. In addition, reducing
shipping distances for material and products to the project minimizes fuel requirements for transportation and
handling. This credit is worth 1 point for the quantities quoted above and 2 points for double the amount, or
20% of the materials.

482

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.11.1 Energy codes

linked to smog, acid rain, and climate change. In the
U.S. most buildings are constructed to meet minimum
energy code requirements; energy codes contribute
to sustainability by saving energy and protecting the
environment.
Energy codes are effective in reducing per capita energy usage (energy use per person). The per capita energy use in California has remained steady due to the
state’s active use and enforcement of energy codes for
buildings, while in the rest of the U.S. that energy use
has increased (Fig. 5.4.10).
The U.S. Energy Conservation and Production Act42 requires that each state certify it has a commercial building code that meets or exceeds ANSI/ASHRAE/IESNA
Standard 90.1. In this sense, “commercial” means all
buildings that are not low-rise residential (three stories
or less above grade). This includes office, industrial,
warehouse, school, religious, dormitories, and highrise residential buildings. The ASHRAE standard and
most codes recognize the benefits of thermal mass and
require less insulation for mass walls.
Thermal mass in exterior walls have the following
benefits and characteristics:

3. Works best in commercial applications.
4. Works well in residential applications.
5. Works best when mass is exposed on the inside
surface.
6. Works well regardless of the placement of mass.
Mass works well in commercial applications by delaying the peak summer load, which generally occurs
around 3:00 p.m. to later when offices begin to close.
As a case in point, the blackout in the northeastern
U.S. in August 2003 occurred at 3:05 p.m.43 A shift
in peak load would have helped alleviate the demand
and, possibly, this peak power problem.
Also, many commercial and industrial customers incur
significant time-of-use utility rate charges for the highest use of electricity for any 1 hour in a month in the
summer. Thermal mass may help shift the peak hour
of electric demand for air conditioning to a later hour,
and help reduce these time-of-use charges. Nighttime
ventilation can be used to cool thermal mass that has
been warmed during the day. Local outdoor humidity
levels influence the effectiveness of nighttime ventilation strategies.

2. Reduces total loads in many climates and locations.

As occupant and equipment heat is generated, it is absorbed not only by the indoor ventilated air but also by
the massive elements of the building. Mass works well

42 1992 National Energy Policy Act, U.S. Department of Energy, www.DOE.
gov.

43 U.S. Department of Energy, Final Report on the August 14, 2003 Blackout
in the United States and Canada: Causes and Recommendations, 2004,
Washington, DC.

1. Delays and reduces peak loads.

5

Fig. 5.4.10
Energy savings due to implementation of energy codes in 1976 in California (California Energy Commission).
Total Electricity Use, per capita, 1960 – 2001
(estimated for California in 2000 and 2001)
kWh

14,000
12,000

12,000

U.S.

8,000

8,000
7,000

6,000
California
4,000
2,000
0

19
60
19
62
19
64
19
66
19
68
19
70
19
72
19
74
19
76
19
78
19
80
19
82
19
84
19
86
19
88
19
90
19
92
19
94
19
96
19
98
20
00

KWh

10,000

ARCHITECTURAL PRECAST CONCRETE

|

483

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.11.1 Energy codes / 5.4.11.4 Advanced energy guidelines

LEED Energy and Atmosphere Prerequisite 2
on Minimum Energy Performance. All buildings
must comply with certain sections on building energy efficiency and performance as required by the
ANSI/ASHRAE/IESNA 90.1-2004, Energy Standard for
Buildings Except Low-Rise Residential Buildings, or
the local energy code, whichever is more stringent.
The ASHRAE standard is usually more stringent and
applies for most states. This prerequisite is a requirement and is not worth any points. The requirements
of the ASHRAE standard are cost-effective and not
particularly stringent for concrete. Insulating to meet
or exceed the requirements of the standard is generally a wise business choice. Determining compliance
for the envelope components is relatively straightforward using the tables in Chapter 5 of the ASHRAE
standard. Minimum requirements are provided for
mass and non-mass components such as walls and
floors.

on the inside surfaces by absorbing the heat gains generated by people and equipment indoors. Interior mass
from interior walls, floors, and ceiling will help moderate
room temperatures and reduce peak energy use.
Thermal mass is most effective in locations and seasons where the daily outdoor temperature rises above
and falls below the balance point temperature of the
building. The balance point temperature is the outdoor
temperature below which heating will be required. It
is less than room temperature, generally between 50
and 60°F (10 and 15°C), at the point where internal
heat gains are about equal to the heat losses through
the building envelope. In many climates, buildings with
thermal mass have lower energy consumption than
non-massive buildings with walls of similar thermal resistance. In addition, heating and cooling needs can be
met with smaller equipment sizes.
More information on thermal properties and energy
code compliance of precast concrete walls is available
in Section 5.3.

5.4.11.2 L ighting
Light-colored precast concrete and other surfaces will
reduce energy costs associated with indoor and outdoor lighting. The more reflective surfaces will reduce
the amount of fixtures and lighting required. Lightcolored precast concrete exposed to the interior will

484

|

ARCHITECTURAL PRECAST CONCRETE

help reduce interior lighting requirements, and lightcolored exterior walls will reduce outdoor lighting
requirements.

5.4.11.3 Air infiltration
Precast concrete panels have negligible air infiltration.
Minimizing air infiltration between panels and at floors
and ceilings will provide a building with low air infiltration. These effects will lower energy costs and help
prevent moisture problems from infiltration of humid
air. In hot and humid climates in the southeastern U.S.,
infiltration of moist air is a source of unsightly and unhealthy moisture problems in buildings. Some building codes44 now limit air leakage of building materials
to 0.004 cfm/ft2 (0.0012 m3/min/m2) under a pressure
differential of 0.3 in. (7.6 mm) water (1.57 psf [0.75
kPa]); precast concrete meets this requirement. Table
5.4.4 lists the measured air leakage values for selected
building materials.

5.4.11.4 Advanced energy guidelines
Sustainability or green building programs (such as
LEED™ or EnergyStar) encourage energy savings beyond
minimum code requirements. The energy saved is a costsavings to the building owner through lower monthly
utility bills and smaller, less expensive heating, ventilat-

Table 5.4.4 Measured Air Leakage for Selected Building Materials.

Material
6 mil (0.15 mm)
polyethylene

Average Leakage at 0.3 in.
Water, cfm/ft2 Surface
No measurable leakage

1 in. (25 mm) expanded
polystyrene

1.0

/2 in. (12 mm)
fiberboard sheathing

0.3

Breather type building
membranes

0.002 – 0.7

1

Closed cell foam
insulation

0.0002

Uncoated brick wall

0.3

Uncoated concrete block

0.4

Precast concrete wall

No measurable leakage

44 Massachusetts Energy Code, www.mass.
gov/bbrs/780_CMR_Chapter_13.pdf.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.11.4 Advanced energy guidelines/ 5.4.12 Indoor Environmental Quality

ing, and air-conditioning (HVAC) equipment. Some government programs offer tax incentives for energy-saving features. Other programs offer reduced mortgage
rates. The EnergyStar program offers simple computer
programs to determine the utility savings and lease upgrades associated with energy saving upgrades.
Many energy-saving measures are cost-effective even
though they exceed minimum codes. Insulation and
other energy-saving measures in building codes generally have a payback of about 5 years, even though
the building life may be anywhere from 30 to 100
years. The New Buildings Institute has developed the
E-Benchmark guidelines to save energy beyond codes
(see www.NewBuildings.org). ASHRAE Advanced
Energy Design Guide For Small Office Buildings (see
www.ASHRAE.org) has a similar purpose. Many utilities are interested in these advanced guidelines to delay the need for new power plants.
The panelized construction of precast concrete lends
itself to good practice and optimization of insulation
levels. To maximize the effectiveness of the insulation,
thermal bridges should be minimized or avoided. Metal
thermal bridges may occur as connectors that penetrate
insulation to connect concrete layers. Concrete thermal
bridges may occur at the bottom and top of concrete
panels. Using fiberglass or carbon-fiber composite fasteners or thermal breaks may minimize thermal bridges.

5

LEED Energy Credit 1 on Optimizing Energy
Performance. This credit is allowed if energy cost savings can be shown compared to a base building that
meets the requirements of ANSI/ASHRAE/IESNA 90.12004, Energy Standard for Buildings Except Low-Rise
Residential Buildings. The method of determining energy
cost savings must meet the requirements of Appendix G
“Performance Rating Method” of the standard.
Many engineering consulting firms have the capability to
model a building to determine energy savings as required
using a computer-based program such as DOE2. When
concrete is considered, it is important to use a program
like DOE21 that calculates annual energy use on an hourly
basis. Such programs are needed to capture the beneficial thermal mass effects of concrete. Insulated concrete
systems, used in conjunction with other energy savings
measures will most likely be eligible for LEED points. The
number of points awarded will depend on the building,
climate, fuel costs, and minimum requirements of the
standard. From 1 to 10 LEED points are awarded for energy cost savings of 10.5% to 42% for new buildings
and 3.5% to 35% for existing buildings (Table 5.4.5). A
small office building less than 20,000 ft2 (1900 m2) complying with ASHRAE “Advanced Energy Design Guide For
Small Office Buildings 2004” can achieve 4 points, and
a building complying with “E-Benchmark” v1.1 (www.
newbuildings.org) can achieve 1 point.
1 Visual DOE 4.0, Architectural Energy Corporation, Boulder, CO, www.
archenergy.com.

Table 5.4.5 LEED NC v2.2 Points Awarded for Energy Costs Saved
Beyond Minimum Code.

New Buildings,
Energy Saved

Existing Buildings,
Energy Saved

Points

10.5%

3.5%

1

14%

7%

2

17.5%

10.5%

3

21%

14%

4

24.5%

17.5%

5

28%

21%

6

31.5%

24.5%

7

35%

28%

8

38.5%

31.5%

9

42%

35%

10

5.4.12 Indoor Environmental Quality
Concrete contains low to negligible VOCs. These
compounds degrade indoor air quality when they off
gas from new products, such as interior finishings, carpet, and furniture. Manufactured wood products such
as laminate, particle board, hardboard siding, and
treated wood can also lead to off gassing. In addition,
VOCs combine with other chemicals in the air to form
ground-level ozone. Table 5.4.6 presents the VOC concentration and emission rates for common materials.
Complaints due to poor indoor air quality routinely include eye, nose, and throat irritation; dryness of the
mucous membranes and skin; nose bleeds; skin rash;
mental fatigue and headache; cough; hoarseness;
wheezing; nausea; dizziness; and increased incidence
of asthma.

ARCHITECTURAL PRECAST CONCRETE

|

485

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.4.12 Indoor Environmental Quality / 5.4.14 Innovation

Polished concrete floors do not require carpeting.
Exposed concrete walls do not require finishing materials–this eliminates particulates from sanding drywall tape
seams. VOCs in concrete construction can be further
reduced by using low-VOC materials for form release
agents, curing compounds, dampproofing materials,
wall and floor coatings and primers, membranes, sealers,
and water repellents.

5.4.13 Demolition
Precast concrete panels can be reused when buildings
are expanded and precast concrete can be recycled as
road base or fill at the end of its useful life. Concrete
pieces from demolished structures can be reused to
protect shorelines. Most concrete from demolition in
urban areas is recycled and not placed in landfills.

Table 5.4.6 Concentrations and Emission Rates of VOCs for Common Materials.

VOC Concentration,
mg/m3

VOC Emission
Rate, mg/m2h

Concrete with waterbased form-release agent

0.018

0.003

Acryl latex paint

2.00

0.43

Epoxy, clear floor varnish

5.45

1.3

Felt carpet

1.95

0.080

Gypsum board

N/A

0.026

Linoleum

5.19

0.22

Building Material

Particle board

N/A

2.0

Plastic silicone sealer

77.9

26.0

Plywood paneling

N/A

1.0

Putty strips

1.38

0.34

PVA glue cement

57.8

10.2

Sheet vinyl flooring

54.8

2.3

Silicone caulk

N/A

<2.0

1,410.0

271.0

Water-based EVA wall
and floor glue

LEED Materials Credit 2 on Construction Waste Management. This
credit is extended for diverting construction and demolition debris and
land clearing waste from landfill disposal. It is awarded based on diverting at
least 50% by weight or volume of the
previously listed materials. Since precast
concrete is a relatively heavy construction material and is frequently crushed
and recycled into aggregate for road
bases or construction fill, this credit
should be obtainable when concrete
buildings are demolished. This credit is
worth 1 point if 50% of the construction, and demolition debris and land
clearing waste is recycled or salvaged
and 2 points for 75%.

5.4.14 Innovation

Note: 1 mg/m3 = 0.000009 oz/yd3; 1 mg/m2h = 0.00001 oz/yd2h.

LEED Indoor Environmental Quality Credit 3.1 on Construction
IAQ Management Plan, During Construction. This credit prevents
indoor air quality problems resulting from the construction process. The
intent is to reduce and contain dust and particulates during construction and to reduce moisture absorbed by materials that are damaged by
moisture. During construction, the project must meet or exceed the recommended Design Approaches of the Sheet Metal and Air Conditioning
National Contractors Association (SMACNA) IAQ Guidelines for Occupied
Buildings under Construction, 1995, Chapter 3 on Control Measures
(www.smacna.org). Using precast concrete can help meet the requirements because it is delivered to the site in pieces that do not require
fabrication, processing, or cutting, thereby reducing dust and airborne
contaminants on the construction site. Concrete is not damaged by
moisture and does not provide nutrients for mold growth. This credit is
worth one point.

486

|

ARCHITECTURAL PRECAST CONCRETE

LEED Innovation and Design Process
Credit 1. This credit is available for
projects that demonstrate exceptional
performance above the requirements
in LEED or not specifically addressed in
LEED. For example, close collaboration
with engineers on a given project to develop innovative systems that are more
resource efficient or use less energy may
earn a project an additional point. To
earn credits (up to 4), the user must submit the intent of the proposed credit, the
proposed requirement for compliance,
submittals to demonstrate that compliance, and the design approach used to
meet the requirement.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.4.15 Conclusion / 5.5.1 Acoustical Properties Glossary

LEED Innovation and Design Process Credit 2.
One point is also given if a principal participant of the
project team is a LEED Accredited Professional. The
concrete industry has LEED-experienced professionals available to assist teams with concrete applications and help maximize points for concrete.

5.4.15 C onclusion
Sustainable practices contribute to saving materials
and energy and reducing the negative effects of pollutants. The use of precast concrete contributes to these
practices by incorporating integrated design, using
materials efficiently, and reducing construction waste,
site disturbance, and noise. Concrete is durable, resistant to corrosion and impact, and inedible.
Precast concrete structures are resistant to fires, wind,
hurricanes, floods, earthquakes, wind-driven rain, blast
forces, and moisture damage. Light- or natural-colored
concrete reduces heat islands, thereby reducing outdoor temperatures, saving energy, and reducing smog.
Recycled materials such as fly ash, slag cement, silica
fume, and recycled aggregates can be incorporated
into concrete, thereby reducing the amount of materials that are taken to landfills and reducing the use of
virgin materials.
Concrete structures in urban areas are recycled into
fill and road base material at the end of their useful life.
Cement and concrete are generally made of abundant
local materials. The thermal mass of concrete helps
save heating and cooling energy in buildings. Concrete
acts as an air barrier, reducing air infiltration and saving more energy. Concrete has low VOC emittance and
does not degrade indoor air quality.
Sustainability attributes can be evaluated by performing a life cycle assessment. Because these procedures
are time consuming, green building rating systems such
as LEED have become popular. Precast concrete can
help a project earn up to 23 points towards LEED certification for new buildings (a total of 26 are required.)

5.5 A
 COUSTICAL PROPERTIES
5.5.1 G
 lossary
Airborne sound – Sound that reaches the point of
interest by propagation through air.

5

Background level – The ambient sound-pressure level
existing in a space
Decibel (dB) – A logarithmic unit of measure of sound
pressure or sound power. Zero on the decibel scale corresponds to a standardized reference pressure (20µPa)
or sound power (10-¹² watt).
Flanking transmission – Transmission of sound by indirect paths other than through the primary barrier.
Frequency (Hz) – The number of complete vibration
cycles per second.
Noise – Unwanted sound.
Noise criteria (NC) – A series of curves, used as design goals to specify satisfactory background sound
levels as they relate to particular use functions.
Noise reduction (NR) – The difference in decibels
between the space-time average sound pressure levels produced in two enclosed spaces by one or more
sound sources in one of them.
Noise reduction coefficient (NRC) – The arithmetic
average of the sound absorption coefficients at 250,
500, 1000, and 2000 Hz expressed to the nearest multiple of 0.05.
RC curves – A revision of the NC curves based on empirical studies of background sounds.
Reverberation – The persistence of sound in an enclosed or partially enclosed space after the source of
sound has stopped.
Sabin – The unit of measure of sound absorption.
Sound absorption coefficient – The fraction of randomly incident sound energy absorbed or otherwise
not reflected off a surface.
Sound pressure level (SPL) – Ten times the common
logarithm of the ratio of the square of the sound pressure to the square of the standard reference pressure
of 20µPa. Commonly measured with a sound level
meter and microphone, this quantity is expressed in
decibels.
Sound transmission class (STC) – The single number
rating system used to give a preliminary estimate of the
sound insulation properties of a partition system. This
rating is derived from measured values of transmission
loss.
Sound transmission loss (TL) –Ten times the com-

ARCHITECTURAL PRECAST CONCRETE

|

487

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.5.1 Acoustical Properties Glossary / 5.5.3 Sound Levels

Structure-borne sound – Sound that reaches the
point of interest over at least part of its path by vibration of a solid structure.

5.5.2 G
 eneral
The basic purpose of architectural acoustics is to provide a satisfactory environment in which the desired
sounds are clearly heard by the intended listeners and
the unwanted sounds (noise) are isolated or absorbed.
The sound-reduction needs of a building are determined based on location, environmental ambiance,
and the degree of sound reduction necessary for occupants to function effectively.
Under most conditions, the architect can design the
building to satisfy the acoustical needs of the tenant.
Good acoustical design uses reflective and absorptive
surfaces, sound barriers, and vibration isolators. Some
surfaces must reflect sound so that the loudness will be
adequate in all areas where listeners are located. Other
surfaces must absorb sound to avoid echoes, sound
distortion, and long reverberation times. Sound is isolated from rooms where it is not wanted by selected
wall and floor/ceiling constructions. Vibrations generated by mechanical equipment are isolated from the
structural frame of the building by means of mechanical isolators or compressible materials.

concrete wall; similarly, low sound transmission is not
available from a porous, lightweight material that may
be applied to room surfaces for sound absorption. It
is important to recognize that the basic mechanisms
of sound absorption and sound insulation are quite
different.

Fig. 5.5.1
Sound transmission class as a function of wall weight.
65

Sound absorbing materials and sound insulating materials are used for two different purposes. There is
not much sound absorption from an 8 in. (200 mm)

488

|

ARCHITECTURAL PRECAST CONCRETE

50
STC = 0.1304 W + 43.48
Statistical Tolerance ± 2.5 STC

45

50

60

70

80

90

100

Weight Per Unit Area - (W) - psf

Fig. 5.5.2
Acoustical test data of solid flat concrete panels –
normalweight concrete.
Sound Transmission Loss
70
8 in. Flat Panel, STC-58
Sound Transmission Loss dB

The problems of sound insulation are usually considerably more complicated than those of sound absorption. Sound insulation involves greater reductions in
sound level than can be achieved by absorption. These
large reductions can only be achieved by continuous,
impervious barriers. If the problem also involves structure-borne sound, it may be necessary to introduce resilient layers or discontinuities into the barrier.

Flat or Ribbed Panels
55

40
40

Most acoustical situations can be described in terms
of: (1) sound source, strength, and path; (2) sound
transmission path; and (3) sound receiver.

5.5.3 S ound Levels

60

60

6 in. Flat Panel, STC-55

50

40

30

4 in. Flat Panel, STC-49

100
125
160
200
250
315
400
500
630
800
1000
1250
1600
2000
2500
3150
4000
5000

mon logarithm of the ratio, expressed in decibels, of
the airborne sound power incident on the partition
that is transmitted by the partition and radiated on the
other side.

Sound Transmission Class (STC)

5

Frequency, HZ

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.5.4 Sound Transmission Loss / 5.5.5 Absorption of Sound

5.5.4 S ound Transmission Loss
Sound transmission loss measurements are made at
16 frequencies at one-third octave intervals covering
the range from 125 to 4000 Hz. The testing procedure
is described in ASTM E 90, Laboratory Measurement of
Airborne Sound Transmission Loss of Building Partitions.
Measurements can also be made in buildings by following ASTM E 336, Measurement of Airborne Sound
Insulation in Buildings. To simplify specification of desired performance characteristics the single number
Sound Transmission Class (STC) (see ASTM E 413) was
developed. It was originally designed to assess sound
Table 5.5.1 Airborne Sound Transmission Class Ratings from
Tests of Precast Concrete Assemblies.

Assembly
No.
Description

STC1
(OITC)

1

4 in. flat panel, 54 psf

49 (43)

2

5 in. flat panel, 60 psf

522

3

6 in. flat panel, 75 psf

55 (46)

4

Assembly 2 with “Z” furring
channels, 1 in. insulation and
1
/2 in. gypsum board, 75.5 psf

62

5

Assembly 2 with wood
furring, 11/2 in. insulation and
1
/2 in. gypsum board, 73 psf

63

6

Assembly 2 with 1/2 in. space,
15/8 in. metal stud row, 11/2
in. insulation and 1/2 in.
gypsum board

632

7

8 in. flat panel, 95 psf

8

10 in. flat panel, 120 psf

58 (50)
592

1 The STC of sandwich panels is about the same as the STC of the thickness
of the two concrete wythes (ignoring the insulation thickness).
2 Estimated values.

Table 5.5.2 Typical Improvements for Wall Treatments Used with
Precast Concrete Elements.

Treatment
Wall furring, 3/4 in. insulation and 1/2 in.
gypsum board attached to concrete wall
Separate metal stud system, 11/2 in.
insulation in stud cavity and 1/2 in. gypsum
board attached to concrete wall
Plaster direct to concrete

Increased
Airborne
STC
3
5 to 10
0

5

(human speech) privacy for interior walls, but its use
has expanded to cover virtually all types of partitions
and partition elements.
Airborne sound reaching a wall, floor, or ceiling produces vibrations in the wall that are radiated
with reduced intensity on the other side. Airborne
sound transmission loss in wall assemblies is a function of their weight, stiffness, and vibration damping
characteristics.
Weight is concrete’s greatest asset when it is used as
a sound insulator. For sections of similar design, but
different weights, the STC increases approximately 6
units for each doubling of weight (Fig. 5.5.1). This figure describes sound transmission class as a function of
weight based on experimental data. Precast concrete
walls usually do not need additional treatments in order to provide adequate sound insulation. If desired,
greater sound insulation can be obtained by using a
resiliently attached layer(s) of gypsum board or other
building material. The increased transmission loss occurs because the energy flow path is increased to include a dissipative air column and additional mass.
The acoustical test results of airborne sound transmission loss of 4, 6, and 8 in. (100, 150, and 200 mm) solid
flat panels are shown in Fig. 5.5.2. Table 5.5.1 presents
the ratings for various precast concrete assemblies. The
effects of various assembly treatments on sound transmission can also be predicted from results of previous
tests shown in Table 5.5.2. The improvements are additive, but in some cases the total effect may be slightly
less than the sum.
The mass of the precast/prestressed concrete loadbearing sandwich wall panels prevented outside noises
from entering the building in Fig. 5.5.3. The design of
this auditorium required selected areas of high resolution and reflectivity, which was achieved by using the
8 in.-thick (200 mm) curved interior wall panels to distribute sound throughout the hall in a geometrically
controlled fashion. They also serve as structural members. Some 200 curved, sandblasted panels, employing
eight different radii, were created to meet all of the
acoustical requirements. They were given a staining
sealer for aesthetic effects.

5.5.5 Absorption of Sound
A sound wave always loses part of its energy as it
is reflected by a surface. This loss of energy is called

ARCHITECTURAL PRECAST CONCRETE

|

489

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.5.5 Absorption of Sound

(b)

Precast concrete controls the acoustics.

Fig. 5.5.3
The Juanita K. Hammons Hall for
the Performing Arts,
Springfield, Missouri;
Architect: Pellham-PhillipsHagerman and Butler, Rosenbury &
Partners (joint venture);
Photos: Pellham-Phillips-Hagerman.

(a)

490

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.5.5 Absorption of Sound/ 5.5.6 Acceptable Noise Criteria

noise levels in specific rooms to assist in evaluating
noise-reduction problems.
The main criticism of NC curves is that they are too
permissive when the control of low or high frequency noise is of concern. For this reason, room criteria
(RC) curves were developed (Fig. 5.5.5). RC curves are
the result of extensive studies based on the human
response to both sound-pressure level and frequency
Fig. 5.5.4 Noise criteria (NC) curves.
90
80

Octave Band Sound Pressure Level, dB re 20 micropascals

sound absorption. It appears as a decrease in sound
pressure of the reflected wave. The sound absorption
coefficient is the fraction of energy incident but not reflected per unit of surface area. Sound absorption can
be specified at individual frequencies or as an average
of absorption coefficients (NRC). A dense, non-porous
concrete surface typically absorbs 1 to 2% of incident
sound and has an NRC of 0.015. In cases where additional sound absorption is desired, a coating of acoustical material can be spray-applied, acoustical tile can
be applied with adhesive, or an acoustical ceiling can
be suspended. Most of the spray-applied fire-retardant
materials used to increase the fire resistance of precast concrete and other floor-ceiling systems can also
be used to absorb sound. The NRC of the sprayed fiber types range from 0.25 to 0.75. Most cementitious
types have an NRC from 0.25 to 0.50.

5

5.5.6 A
 cceptable Noise Criteria
As a rule, a certain amount of continuous sound can
be tolerated before it becomes noise. An “acceptable”
level neither disturbs room occupants nor interferes
with the communication of wanted sound.
The most generally accepted noise criteria (NC) used
today are expressed as the Noise Criteria or the Room
Criteria (RC) curves (Fig. 5.5.4, Table 5.5.3 and Fig.
5.5.5).
The figures in Table 5.5.4 represent general acoustical goals. They can also be compared with anticipated

70
NC-65

60

NC-60
NC-55

50

NC-50
NC-45

40

NC-40
NC-35

30

NC-30

Approximate
20 threshold of
hearing for
continuous
10 noise
16 125 250

NC-25
NC-20
NC-15
500 1000 2000 4000 8000

Octave Band Center Frequencies, Hz

Table 5.5.3 Data for noise criteria curves.

Octave Band Center Frequency, Hz

Noise Criteria
Curves

63

125

250

500

1000

2000

4000

8000

NC-15

1

47

36

29

22

17

14

12

11

NC-20

1

51

40

33

26

22

19

17

16

NC-25

1

54

44

37

31

27

24

22

21

NC-30

57

48

41

35

31

29

28

27

NC-35

60

52

45

40

36

34

33

32

NC-40

64

56

50

45

41

39

38

37

NC-45

67

60

54

49

46

44

43

42

NC-50

71

64

58

54

51

49

48

47

NC-55

74

67

62

58

56

54

53

52

NC-60

77

71

67

63

61

59

58

57

NC-65

80

75

71

68

66

64

63

62

1 The applications requiring background levels less than NC-25 are special purpose spaces in which an
acoustical consultant should set the criteria.


ARCHITECTURAL PRECAST CONCRETE

|

491

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.5.6 Acceptable Noise Criteria

and take into account the requirements for speech
intelligibility.

Fig. 5.5.5 Room criteria (RC) curves.

Table 5.5.4 Recommended Category Classification and Suggested
Noise Criteria Range for Steady Background Noise as Heard in
Various Indoor Functional Activity Areas.1

Type of Space

NC or RC
Curve

1. Private residence

25 to 30

2. Apartments

30 to 35

3. Hotels/motels
a. Individual rooms or suites

30 to 35

b. Meeting/banquet rooms

30 to 35

c. Halls, corridors, lobbies

35 to 40

d. Services/support areas

40 to 45

4. Offices
a. Executive

25 to 30

b. Conference rooms

25 to 30

c. Private

30 to 35

d. Open-Plan areas

35 to 40

e. Computer/business machine areas

40 to 45

f. Public circulation

40 to 45

5. Hospitals and clinics
a. Private rooms

25 to 30

b. Wards

30 to 35

c. Operating rooms

25 to 30

d. Laboratories

30 to 35

e. Corridors

30 to 35

f. Public areas

35 to 40

6. Churches

25 to 302

7. Schools
a. Lecture and classrooms

25 to 30

b. Open-Plan classrooms

30 to 352

8. Libraries

30 to 35

9. Concert halls

2

10. Legitimate theaters2
11. Recording studios2
12. Movie theaters

30 to 35

1 Design goals can be increased by 5dB when dictated by budget constraints
or when noise intrusion from other sources represents a limiting condition.
2 An acoustical expert should be consulted for guidance on these critical spaces.

492

|

ARCHITECTURAL PRECAST CONCRETE

Octave Band Sound Pressure Level, dB re 20 micropascals

90
A

80

B
70
60
50
40

RC
50
45

C

40

30

35
30

20
10
16

25
31.5

63

125 250

500 1000 2000 4000

Octave Band Center Frequencies, Hz
Region A: H
 igh probability that noise-induced vibration levels in
lightweight wall/ceiling constructions will be clearly
perceptible; anticipate audible rattles in low mass
fixtures, doors, windows, etc.
Region B: N
 oise-induced vibration levels in lightweight wall/ceiling
constructions may be moderately perceptible; slight possibility of rattles in low mass fixtures, doors, windows, etc.
Region C: B
 elow threshold of hearing for continuous noise.

A low background level obviously is necessary
where listening and speech intelligibility is important.
Conversely, higher ambient levels can persist in large
business offices or factories where speech communication is limited to short distances. Often, the minimum
target levels are just as important as the maximum permissible levels listed in Table 5.5.4. In an office or residence, it is desirable to have a certain ambient sound
level to assure adequate acoustical privacy between
spaces and minimize the transmission loss requirements of unwanted sound (noise).
These undesirable sounds may be from exterior
sources such as automobiles and aircraft, or they may
be generated as speech in an adjacent classroom or
music in an adjacent apartment. They may also be
direct impact-induced sound such as footfalls on the
floor above, rain on a lightweight roof construction,
or vibrating mechanical equipment. Thus, the designer
must always be ready to accept the task of analyzing
the many potential sources of intruding sound as related to their frequency characteristics and the rates at
which they occur. The level of toleration that is to be

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.5.6 Acceptable Noise Criteria/ 5.5.7 Composite Wall Considerations

expected by those who will occupy the space must also
be established. Figures 5.5.6 and 5.5.7 are the spectral
characteristics of common noise sources.
With these criteria, the problem of sound isolation
now must be solved, namely the reduction process
between the high, unwanted noise source and the
desired ambient level. Once the objectives are estabFig. 5.5.6 Sound pressure levels — exterior noise sources.
120

Jet Aircraft Takeoff
500 ft

100

Sound Pressure Level, dB

Bus
Propeller Aircraft
Takeoff 500 ft

80
Heavy Truck – 20 ft
60

Automobiles – 20 ft

20

63

125

250

500

lished, the designer then should refer to available data
(for example in Fig. 5.5.1 or Table 5.5.1) and select
the system that best meets these requirements. In this
respect, precast concrete systems have superior properties and can, with minimal effort, comply with these
criteria. When the insulation value has not been specified, selection of the necessary barrier can be determined analytically by (a) identifying exterior and /or interior noise sources, and (b) by establishing acceptable
interior noise criteria.
Example: Sound Insulation Criteria
Assume a precast concrete office building is to be
erected adjacent to a major highway. Private and semiprivate offices will run along the perimeter of the structure. The first step is to determine the degree of insulation required of the exterior wall system (see Sound
Pressure Level 1). The NC data is used because it is
more familar to and preferred by designers.
The 500 Hz requirement, 38 dB, can be used as the
first approximation of the wall STC category. However,
if windows are planned for the wall, a system of about
50–55 STC should be selected (see following composite wall discussion). Individual transmission loss performance values of this system are then compared to the
calculated need (see Sound Pressure Level 2).

40

0

5

1000

2000 4000

8000

Frequency, Hz

Fig. 5.5.7 Sound pressure levels — interior noise sources.

The selected wall should meet or exceed the insulation needs at all frequencies. However, to achieve the
most efficient design conditions, certain limited deficiencies can be tolerated. Experience has shown that
the maximum deficiencies are 3 dB at two frequencies
or 5 dB on one frequency point.

120

5.5.7 Composite Wall Considerations

Riveting

Sound Pressure Level, dB

100

Stereo Phnograph,
Teenager Lever

Typical Office

80
Business Machine
Tabulating Room

60

Bed or Dining Room
40

Kitchen

20

0

63

125

250

500

1000

Frequency, Hz

2000 4000

8000

An acoustically composite wall is made up of elements of varying acoustical properties. Windows and
doors are often the weak link in an otherwise effective
sound barrier. Minimal effects on sound transmission
loss will be achieved in most cases by proper selection
of glass (Table 5.5.5). The control of sound transmission through windows requires large cavities between
layers (multiple glazing), heavy layers (thicker glass),
laminated glass, and reduction of the structural connection between layers (separate frames and sashes
for inner and outer layers). Also, mounting of glass
lites with soft neoprene edge gaskets may not be as
effective at reducing sound transmission as systems
that use wet seals (gunable sealants). The combination
of wet seals with butyl tape or open cell foam dra-

ARCHITECTURAL PRECAST CONCRETE

|

493

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.5.7 Composite Wall Considerations

Sound Pressure Level 1.

Sound Pressure Level – (dB)
Frequency (Hz)

63

125 250 500 1000 2000 4000 8000

Bus traffic source
noise (Fig. 5.5.6)

80

83

85

78

74

68

62

58

Private office noise
criteria – NC 35
(Fig. 5.5.4)

60

52

45

40

36

34

33

32

Required insulation

20

31

40

38

38

34

29

26
Sound Pressure Level 2.

matically reduces the potential for air infiltration,
and therefore, flanking sound transmission. They
certainly have to be as airtight as possible; usually
fixed windows provide much better sound transmission control than operable windows.

Sound Pressure Level – (dB)

Sound pressure impinging on the window framing will cause it to vibrate, transmitting sound to
the building interior. Consequently, the window-

Frequency (Hz)

125 250 500 1000 2000 4000

Required insulation

31

40

38

38

34

29

6 in. precast
concrete solid
concrete wall
(Fig. 5.5.2)

38

43

52

59

67

72

-

-

-

-

-

-

Deficiencies
Table 5.5.5 Acoustical Properties of Glass.

Sound Transmission Class (STC)
Type and Overall
Thickness, in.

Inside Lite, in.

Outside Lite, in.

STC

OITC

/8

1

/8

31

26

—

—

1

/4

31

29

—

5

/8 Insulated Glass

1

1

/4 Plate or Float

1

/2 Plate or Float

1 Insulated glass
1

—

/2

36

32

/4

1

/2 Air space

1

/4

35

28-30

/8

0.030 Vinyl

1

/8

35

—

/16 Air space

3

/16

37

28-30

—

3

/2 Air space

1

1

1 /2 Insulated glass

/4

1

3

/8

1

/4 Plate or Float

9

—

1 Insulated glass

1

3

1

1

/4 Laminated

Construction
Space, in.

/4 Laminated

1

1 Plate or Float

—

—

2 /4 Insulated glass

1

/4

2 Air space

1 Laminated Insulated glass

1

/4

3

/4

36

—

/4

39

31

1

37

—

39

—

41

32

/2

1

/2 Air space

/8 plus /8

1

1

1

Transmission loss (dB)
Frequency (Hz)
125

160

200

250

315

400

500
1

25

25

24

28

26

29

630

800

1000

1250

1600

2000

2500

3150

4000

34

30

27

32

37

37

37

43

46

43

49

54

/4 in. plate glass – 31 STC; 29 OITC
31

33

34

34

35

1 in. insulating glass with /2 in. air space – 35 STC; 28 OITC
1

24

29

22

22

25

30

33

35

38

40

42

42

1 in. insulating glass laminated with /2 in. air space – 39 STC; 31 OITC
1

17

494

|

28

29

33

34

38

40

ARCHITECTURAL PRECAST CONCRETE

40

41

41

41

41

40

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.5.7 Composite Wall Considerations

STC is not necessarily the best performance specification for windows as it is often a poor predictor of
sound insulation for low frequency sources, such as
mechanical system or transportation noise. The OITC
(Outdoor-Indoor Transmission Class) rating system
based on ASTM E 1332 is relatively new, and it was
designed to assess a building façade element, such
as a window, when exposed to a standard spectrum
of low frequency air and truck transportation noise
ranging from 80 to 4000 Hz (see ASTM Guide E 966).
Therefore, it is a better measure of a window system’s
performance than STC, especially when traffic noise is
the principal concern. The numeric value representation of OITC tends to be lower than the STC rating.
There are many options available for acoustical glazing, so it is important to make the right choice—especially if the building is exposed to significant exterior
noise and the interior spaces are noise sensitive. The
use of double-pane insulating glass is not adequate
for many projects. Even single- or double-laminated
insulating glass may not be adequate, especially at low
outside temperatures, where regular PVB-laminated
glass will yield a performance similar to that of nonlaminated glass.
The sound-transmission loss through a door depends
on the material and construction of the door and the
effectiveness of the seal between the door and its
frame. There is a mass law dependence of STC on
weight (psf) for both wood and steel doors. The approximate relationships are:
For steel doors: STC = 15 + 27 log W
For wood doors: STC = 12 + 32 log W
where W = weight of the door, psf.
These relationships are purely empirical and a large
deviation can be expected for any given door. ASTM E
1408 can be used to determine the acoustical performance of doors.
For best results, the distances between adjacent
door and/or window openings should be maximized,

staggered when possible, and held to a minimum
area. Minimizing openings allows the wall to retain
the acoustical properties of the precast concrete. The
design characteristics of the door or window systems
must be analyzed prior to specification. Such qualities
as frame design, door construction, and glazing thickness are vital performance criteria. Installation procedures must be exact and care should be given to the
framing of each opening. Gaskets, weatherstripping,
and raised thresholds serve as both thermal and acoustical seals and are recommended.
Figure 5.5.8 can be used to calculate the effective
acoustic isolation of a wall system that contains a composite of elements, each with known individual transmission loss data (TL). (For purposes of approximation,
STC values can be used in place of TL values.)
Example: Composite Wall Insulation Criteria
To complete the office building wall acoustical design
from Section 5.5.6 assume the following:
1. The glazing area represents 10% of the exterior
wall area.
2. The windows will be double glazed with a 40 STC
acoustical insulation rating.
The problem now becomes the test of determining
the combined effect of the concrete-glass combina-

Fig. 5.5.8 Chart for calculating the effective transmission
loss of a composite barrier.
1

TL (Wall) – TL (Door, Window or Opening)

glass performance cannot solely be relied on to reduce
sound transmission to the building interior. The sound
transmission of the window framing will result in higher levels of sound transmission through the glass and
wall. Also, window-framing systems that allow greater
amounts of air infiltration also allow greater sound
transmission.

5

2

10
0
50

5

20
10
5
2
1
0.5
0.2
0.1

10
15
20
30
40
50
60

Percent of total area of
wall occupied by door,
window or opening

1

2

3

4 5 6 7 8 910 15 20

30 40 50 60

Decibels to be Subtracted from TL of Wall for
Effective TL of Composite Barrier

ARCHITECTURAL PRECAST CONCRETE

|

495

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.5.7 Composite Wall Considerations/5.5.8 Leaks and Flanking

Sound Pressure Level 3.

5.5.8 Leaks and Flanking

Sound Pressure Level – (dB)
Frequency (Hz)

125 250 500 1000 2000 4000

6 in. precast solid
concrete wall (Fig. 5.5.2)

38

43

52

59

67

72

Double-glazed windows
(Table 5.5.5)

17

33

40

41

40

54

Correction (Fig. 5.5.8)

10

3

4

9

16

9

Combined transmission
loss

28

40

48

50

51

63

Insulation requirements

31

40

38

38

34

29

Deficiencies

-3

—

—

—

—

—

Note: 1 in. = 25.4 mm

tion and a re-determination of criteria compliance (see
Sound Pressure Level 3).
The maximum deficiency is 3 dB and occurs at only
one frequency point. The 6 in. (150 mm) precast concrete wall with double-glazed windows will provide
the required acoustical insulation.
Floor-ceiling assembly acoustical insulation requirements are determined in the same manner as walls by
using Fig. 5.5.2 and 5.5.8.

Fig. 5.5.9 Effect of safing insulation seals.
6 in. Concrete Floor 55 STC
Sealant

Inorganic Mineral Wool
Insulation

≥ 1/8 in. Steel Bent Plate
Exterior Wall

Combined
Trasmission Loss
No closure

14 STC

With steel bent plate closure

28 STC

With 4 in. thick safing insulation

30 STC

steel bent plate added
With 6 in. thick safing insulation

496

|

Anticipation and prevention of leaks begins at the design stage. Flanking paths (gaps) at the perimeters of
interior precast concrete walls and floors are generally
sealed during construction with grout or drypack. All
openings around penetrations through walls or floors
should be as small as possible and must be sealed airtight. The higher the required STC of the barrier, the
greater the importance of sealing all openings.
Perimeter leakage commonly occurs at the intersection between an exterior cladding panel and a floor
slab. It is of vital importance to seal this gap to retain
the acoustical integrity of the system and provide the
required fire stop between floors. One way to seal the
gap is to place a 4 pcf (64 kg/m3) density mineral wool
blanket between the floor slab and the exterior wall.
Figure 5.5.9 demonstrates the acoustical isolation effects of this treatment. An enhancement to Fig. 5.5.9
would be to recess the insulation below the floor plane
and fill the recess with smoke stop elastomeric sealant.
Thereby improving not only the sound but the snoke
resistance of the assembly.
Flanking paths can be minimized by:

Gap

steel bent plate added

Performance of a building section with an otherwise
adequate STC can be seriously reduced by a relatively
small hole (or any other path) that allows sound to
bypass the acoustical barrier. All noise that reaches a
space by paths other than through the primary barrier
is called flanking noise. Common flanking paths are
openings around doors or windows, electrical outlets,
telephone and television connections, and pipe and
duct penetrations. Suspended ceilings in rooms where
walls do not extend from the ceiling to the roof or floor
above also allow sound to travel to adjacent rooms by
flanking.

42 STC
38 STC
45 STC

ARCHITECTURAL PRECAST CONCRETE

1. Interrupting the continuous flow of energy with
dissimilar materials, that is, expansion or control
joints or air gaps.
2. Increasing the resistance to energy flow with floating floor systems, full height and/or double partitions, and suspended ceilings.
3. Using primary barriers, which are less subject to
the creation of flanking paths. Although not easily
quantified, an inverse relationship exists between
the performance of an element as a primary barrier and its propensity to transmit flanking sound.
In other words, the probability of existing flanking

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.5.8 Leaks and Flanking/ 5.6.1 Design Considerations For Blast Resistance General

The goal of protective design against the effects of
blast is the protection of the building occupants and
the reduction of casualties. Economically feasible design for antiterrorism/force protection (AT/FP) requires
an integrated approach to facility siting, operation programming of interior spaces, and employment of active
and passive security measures employing both technological and human security provisions. The architectural
façade is one element of the protective design chain.
Designing a structure that could face a threat from a
terrorist bombing that could originate either externally
or internally to the structure requires finding the most
effective way to meet the standards for enhanced safety
that currently exist. This section only addresses external
blasts. When designing protection for a building, owners and architects must work with structural engineers
and blast consultants to determine the blast forces to
withstand, as well as risk and vulnerability assessment,
and protection levels. Optimally, blast mitigation provisions for a new building should be addressed in the
early stages of project design to minimize the impact
on architecture and cost. Defensive design often affects aesthetics, accessibility, fire safety regulations,
and budgetary constraints.
The building’s façade is its first real defense against
the effects of an explosion. How the façade responds
to blast loading will significantly affect the behavior
of the structure. The need for comprehensive protection of occupants within the structure will likely cause

Several types of hazards can affect the building systems (structural or architectural). These hazards can
be subdivided into two general categories: man-made
(blast) and natural (earthquakes, wind, etc). For a successful approach to any system design, it is essential to
understand the nature of the hazard. Dynamic hazards
can be described by their relative amplitudes and relative time (frequency) attributes. Figure 5.6.1 shows a
schematic representation of the amplitude-frequency
relationships of several dynamic hazards.
It is important to emphasize the principal differences
between static, dynamic, and short-duration dynamic
loads. Typically, static loads do not produce inertia effects in the structural response, are not time dependent, and are assumed to act on the structure for long
periods of time (gravity loads, for example). Dynamic
loads, such as those induced by earthquake or wind
gusts, have strong time dependencies and their typical
durations are measured in tenths of seconds. Short-

Fig. 5.6.1 Qualitative amplitude-frequency distribution
for different hazards.
High

In today’s environment of enhanced risk some facilities require protective design and the management of
risk of intentional and accidental explosions. There are
many design options available to reduce the risk to any
building.

Medium

5.6 D
 ESIGN CONSIDERATIONS FOR
BLAST RESISTANCE
5.6.1 G
 eneral

Blast
Wind

Seismic
Machine Vibration

Acoustic

Low

If the acoustical design is balanced, the maximum
amount of acoustic energy reaching a space via flanking should not equal the energy transmitted through
the primary barriers. In exterior walls, the proper application of sealant and backup materials in the joints
between units will not allow sound to flank the wall.

window sizes to decrease in height and width, and increase in thickness. Attachments to windows and the
panels themselves likewise will become more substantial. Considering the extent of surface area enclosing a
building, even modest levels of protection will be expensive. As a result, the design philosophy might best
be served by concentrating on the improvement of the
post-damaged behavior of the façade. To protect the
occupants to the highest degree, the aim should be for
the building and its cladding components to remain
standing or attached long enough to protect occupants from injury or death resulting from flying debris
and to evacuate everyone safely.

Amplitude Scale

paths in a concrete structure is much less than in a
structure with steel or wood framing.

5

Very Low

Low

Medium

High

Very High

Frequency Scale
SOURCE: Ettouney, M., “Is Seismic Design Adequate for Blast?” Society of
American military Engineers National Symposium on Comprehensive Force
Protection, Charleston, S.C., November 2001.

ARCHITECTURAL PRECAST CONCRETE

|

497

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.1 Design Considerations For Blast Resistance General / 5.6.2 Blast Basics

duration dynamic loads, such as those induced by
explosions or debris impact, are non-oscillatory pulse
loads, and their duration is about 1000 times shorter
than the duration of typical earthquakes. Structural
response under short-duration dynamic effects could
be significantly different than the much slower loading cases, requiring the designer to provide suitable
structural details. Therefore, the designer must explicitly address the effects related to such severe loading
conditions, in addition to the general principles used
for structural design to resist conventional loads. As a
starting point, the reader should review background
material on structural considerations and design in the
references in the Blast Analyses Standards section.
There are conflicting hazard demands on cladding
relating to the weight or mass of a typical wall. For a
seismic hazard, the forces on the wall are directly proportional to its mass. Forces that are produced from a
blast hazard are more generally inversely proportionate
to the mass of the cladding. In some panel configurations, increasing the mass of the panel can provide
improvements in the response of the panel to a defined level of blast loading. This produces a dilemma
for the designer: higher mass would be beneficial in a
blast condition, but be harmful in an earthquake condition. Obviously, an optimization or balanced design
is needed in such a situation, with the understanding
that both hazards require ductile behavior from the
cladding and connections. However, the manner the
cladding-structure interacts when subjected to each
of the two hazards is completely different. During
earthquakes, the movement of the structure will im-

Fig. 5.6.2 Qualitative pressure-time history.

Pressure

Reflected
Pressure

Incident
Overpressure

Ambient
Pressure

Positive
Impulse Negative
Impulse
Negative
Arrival Positive
Phase
Time Phase
Time After Explosion

SOURCE: “Structures to Resist the Effects of Accidental Explosions,”
TM 5-1300. November, 1990.

498

|

ARCHITECTURAL PRECAST CONCRETE

pose forces on the cladding. During a blast event, the
cladding would impose reactions (through the connections) on the structure.

5.6.2 Blast Basics
An explosion is a very rapid release of stored energy
characterized by an audible blast. Part of the energy is
released as thermal radiation, and part is coupled into
the air (air-blast) and soil (ground-shock) as radially expanding shock waves. Air-blast is the principal damage
mechanism. Air-blast phenomena occur within milliseconds and the local effects of the blast are often over
before the building structure can globally react to the
effects of the blast. Also, initial peak pressure intensity
(referred to as overpressure) may be several orders of
magnitude higher than ambient atmospheric pressure.
The overpressure radiates from the point of detonation
but decays exponentially with distance from the source
and time and eventually becomes negative (outwardrushing force), subjecting the building surfaces to suction forces as a vacuum is created by the shock wave
(Fig. 5.6.2). In many cases, the effect of the negative
phase is ignored because it usually has little effect on
the maximum response. The maximum impulse delivered to the structure is the area under the positive
phase of the reflected pressure-time curve. Both the
pressure and impulse (or duration time) are required to
define the blast loading.
The shape of the building can affect the overall damage to the structure. For example, U- or L-shaped buildings may trap the shock wave, which may increase blast
pressure locally because of the complex reflections created. Large or gradual re-entrant corners have less effect than small or sharp re-entrant corners. In general,
convex rather than concave shapes are preferred for
the exterior of the building. The reflected pressure on
the surface of a circular building is less intense than on
a flat building. The extent of damage depends on the
yield or charge weight (measured in equivalent pounds
of TNT), the relative position of the explosive device,
and the design details. The shock waves compress air
molecules in its path, producing overpressure. When
the shock waves encounter the building surfaces, they
are reflected, amplifying the overpressure so that it is
higher than the initial peak pressure. Blast load pressures can greatly exceed wind and seismic design loads.
Therefore, it is typically costly for these buildings to be
designed to withstand a large explosion in or very near
the building.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.2 Blast Basics / 5.6.3 Blast Analyses Standards

A secondary effect of the air blast is dynamic pressure
or drag loading, which is a very high velocity wind.
It propels the debris generated by the air blast, creating secondary projectiles. Also, the building is subject
to the ground-shock, which produces ground motions
sometimes similar to a short duration earthquake.
The response of a building to a large explosion occurs
in distinct phases. Initially, as the blast wave contacts
the nearest exterior wall of the building, windows may
be shattered, and the walls and columns deflect under
the reflected pressure. If the blast intensity is greater
than that designed for, the wall eventually deforms
inelastically and suffers permanent displacement or
collapse. The internal pressure exerts a downward and
upward pressure on the floor slabs, depending on the
expected performance of the façade in the blast. If the
façade remains intact during a blast event this limits
the propagation of the blast pressures within the building. The upward pressure on the building floor is important because columns and slabs are not ordinarily
designed for such loads. As the blast wave expands
and diffracts around the building, it exerts an overpressure on the roof, side walls and, finally, on the walls of
the far side (Fig. 5.6.3). Although the pressure levels on
the three sides facing away from the blast are smaller
than those on the front, they are significant. Since the
location of the explosion cannot be anticipated, each
building face must be designed for the worst case, that
is, an explosion normal to that face. Internal pressure
may be reduced by decreasing the size and number of
openings or by using blast resistant glazing assemblies
and doors.

Fig. 5.6.3 Blast loading.
Over Pressure
Spherical
Shock
Wave

Over Pressure

Reflected
Pressure Ground
Shock
Drag
Over Pressure

Stand-off

Center
of Burst

Perimeter Protection
(Fence – Guards – Barriers)
SOURCE: “Structures to Resist the Effects of Accidental Explosions,”
TM5-1300. November, 1990.

5

Blast characteristics are very different in open air versus confined spaces. For example, parking structures
have varying degrees of openness or vent area and the
blast response will be very structure-specific. Confined
and contained explosions produce very complex pressures within and exiting from the structure. Confined
explosions include a reflected shockwave phase. The reflected shockwave phase is similar to an open-air blast
except that it is much more complex due to reverberation off of various surfaces in the structure. A second
loading phase is a quasi-static pressure pulse caused
by the overpressure settling to a slowly decaying level.
The overpressure is dependent on the charge-weightto-enclosing room-volume ratio for its peak value and
depends on the vented area of the enclosure for its decay characteristics. The result of these phases is a much
longer lasting and potentially more damaging pressure
being applied to the structure.

5.6.3 Blast Analyses Standards
All building components requiring blast resistance
should meet the criteria required for GSA or DOD
facilities and be designed using established methods
and approaches for determining dynamic loads and
dynamic structural response. Design and analysis approaches should be consistent with those in the technical manuals below:
1. U.S. Departments of the Army, Navy, and Air Force,
“Structures to Resist the Effects of Accidental
Explosions,” Revision 1 (Department of the Army
Technical Manual TM 5-1300, Department of the
Navy Publication NAVFAC P-397, Department of
the Air Force Manual AFM 88-22), Washington,
DC, November 1990. A CD version of this manual
can be obtained from the Department of Defense
Explosive Safety Board (Phone: 703-325-8624).
Hard copies should be requested from the Defense
Technical Documentation Center. (This reference,
in combination with Con Wep software, guides
designers in the calculation of the pressure and related information necessary to perform an analysis
for the structure.) Contact David Hyde, U.S. Army
Engineer Research and Development Center, 3909
Halls Ferry Road, Vicksburg, Mississippi 39180, or
via email at HydeD@wes.army.mil.
2. Conventional Weapons Effects (CONWEP);
Request through U.S. Army Engineer Research
and Development Center, 3909 Halls Ferry Road,
Vicksburg, MS 39180; Contact David Hyde, email

ARCHITECTURAL PRECAST CONCRETE

|

499

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.3 Blast Analyses Standards / 5.6.4 Determination of Blast Loading

at HydeD@wes.army.mil. Restricted distribution.
In general, federal contractors should have their
government program officer contact Mr. Hyde.
3. UFC 3-340-01, June 2002, “Design and Analysis
of Hardened Structures to Conventional Weapon
Effects (DAHS);” (supersedes Army TM 5-8551/ AFPAM 32-1147(I)/ NAVFAC P-1080, and
DAHSCWEMAN-97). For official use only and may
be obtained by government contractors through
government program officers.
4. Unified Facilities Criteria (UFC) 4-010-10, DoD
Minimum Antiterrorism Standards For Buildings.
This is for official use only and may be obtained
by government contractors through a government
program officer. This document will provide design
charge weights for specific building categories and
applicable level of protection.
5. U.S. Department of the Army, Security Engineering,
TM 5-853 and Air Force AFMAN 32-1071, Volumes
1, 2, 3, and 4; Washington, DC, Departments of
the Army and Air Force, 1994.
6. Air Force Engineering and Services Center, “Protective
Construction Design Manual,” ESL-TR-87-57.
Prepared for Engineering and Services Laboratory,
Tyndall Air Force Base, FL., November 1989.
7. U.S. Department of Energy, “A Manual for
the Prediction of Blast and Fragment Loadings
on Structures,” Revision 1, DOE/TIC 11268.
Washington, DC, Headquarters U.S. Department of
Energy, July 1992. The manual is published by the
U.S. Department of Energy, Albuquerque Operations
Office, and is available from the Department of
Defense Technical Information Center.
8. Unified Facilities Criteria, DoD Minimum
Antiterrorism Standards for Buildings, UFC 4-01001, U.S. Department of Defense, October 8, 2003.
To obtain a copy go to www.wbdg.org/references/
pa_dod.php and click on “UFC Documents.” To
insure they are current, copies are distributed only
in electronic media versions.
9. Interim Antiterrorism/Force Protection Construction
Standards—Guidance on Structural Requirements
(DRAFT), U.S. Department of Defense, March 5,
2001.
It is likely that to design against blast will require
a comprehensive knowledge of explosive effects
and fortification sciences, such as described in the
DAHSCWEMAN (1998), in Technical Manual (TM)

500

|

ARCHITECTURAL PRECAST CONCRETE

5-855-1 (U.S. Department of the Army 1998), and in
the Tri-Service manual (TM 5-1300, U.S. Departments
of the Army, Navy, and Air Force, 1990). The electronic
version of the DAHSCWEMAN manual will greatly assist designers in applying blast design concepts.
The report “Structural Design for Physical Security:
State of the Practice,” prepared by the Structural
Engineering Institute Task Committee, Edward J.
Conrath, et al, American Society of Civil Engineers
(1999), also addresses the design of structures to resist
the effects of terrorist bombings. It provides guidance
for structural engineers charged with designing for
blast resistance of civil facilities.

5.6.4 Determination of Blast Loading
Currently there are no formal blast performance
criteria for civilian buildings. The U.S. Department of
Defense, Department of State, and General Services
Administration have developed specific antiterrorism
requirements for military, embassy, and federal buildings, respectively. However, for security reasons, key
portions of these criteria are only available to designers of specific projects to which they apply. Table 5.6.1
provides some recommendations for private-sector
facilities. In all cases the designer’s goal is to balance
the nature and probability of each threat with the additional costs of protecting against it.
The key aspect of structural design to resist blast effects and progressive collapse is determining the nature and magnitude of the blast loading. This involves
assessing the amount and type of explosive as well as
its distance from the building. Another factor is the
level of security that can be placed around the building
to prevent or mitigate exposure to an explosive event.
The design vehicle weapon size that is considered
will usually be much smaller than the largest credible
threat, measured in the hundreds of pounds rather
than the thousands of pounds of TNT equivalent. The
decision regarding the blast design criteria for a particular building is usually based on a trade-off between
the largest credible attack directed against the building
and the design constraints of the project. Further, the
design pressures and impulses may be less than the
actual peak pressures and impulses may be less than
the actual peak pressures and impulses acting on the
building. This is the general approach that the federal
government has taken in its design criteria for federally
owned domestic office buildings.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.4 Determination of Blast Loading/ 5.6.5 Blast Effects Predictions

5

Table 5.6.1 Recommended Antiterrorism Design Criteria (Conrath et. al.).

Tactic

Parameter

Estimated Likelihood of Terrorist
Attack

Measurement of Standoff Distance R

Low

Medium

High

Very High

Vehicle Size*
(lbs GVW)

4,000

4,000

5,000

12,000

Charge Size
W (lbs TNT)

50

100

500

2,000

Placed
Bomb

Charge Size
W (lbs TNT)

0

2

100

100

Unobstructed Space or Unsecured
Parking / Road

Standoff
Weapon

Charge Size
W (lbs TNT)

2

2

50

50

Neighboring Structure

Vehicle
Bomb

Controlled Perimeter, Vehicle Barrier, or
Unsecured Parking/Road

* For barrier design, with maximum velocity based on site configuration.
SOURCE: Schmidt, Jon A., “Structural Design for External Terrorist Bomb Attacks,” NCSEA, Structure magazine (www.structuremag.org), March, 2003.

The total dynamic pressure (in psi) and the positive
phase duration (in milliseconds) are found using TNT
equivalents (the equivalent weight of the explosive in
TNT = W) and the distance from the blast (R). To calculate blast loads, the blast must be scaled. Similar blast
waves are produced at identical scaled distances when
two explosive charges of similar geometry and of the
same explosive, but of different sizes, are detonated in
the same atmosphere. The scaled distance parameter
R
Z (ft per lb TNT equivalent) is: Z = 1/ 3
W
With the scaled distance in the correct units, published curves can be used to find the total dynamic
pressure and the positive phase duration.
Although the angle of incidence at which a blast
wave strikes the building surface also influences these
parameters, it is usually conservative to neglect this adjustment. Either way, in order to obtain the blast load,
a number of different tools can be used. Tables of
pre-determined values may be used (see GSA Security
Reference Manual: Part 3 – Blast Design & Assessment
Guidelines, July 31 2001) or computer programs can
perform these calculations and provide much greater
accuracy. One such software product, AT-Blast, is available for downloading free of charge from the U.S.
General Services Administration (www.oca.gsa.gov
or www.araseas.com). Designers of government projects may request Con Wep, another software product,
through the agency that they have a contract with.
Con Wep is a collection of conventional weapons effects calculations from the equations and curves of TM
5-855-1. Users should be thoroughly familiar with TM
5-855-1 before using this program as a design tool.

Although the actual blast load on an exposed element will vary over its tributary area, for design the
maximum dynamic load is typically taken as the product of this area and either the maximum pressure or a
spatially averaged value. This is analogous to the manner in which design wind loads for components and
cladding are routinely calculated. Blast loads need not
be factored since they already represent an ultimate
design condition.

5.6.5 Blast Effects Predictions
After the blast load has been predicted, damage levels may be evaluated by explosive testing, engineering analysis, or both. Often, testing is too expensive an
option for the design community and an engineering
analysis is performed instead. To accurately represent
the response of an explosive event, the analysis needs
to be time dependent and account for non-linear
behavior.
Non-linear dynamic analysis techniques are similar
to those currently used in advanced seismic analysis.
Analytical models range from equivalent single-degree-of-freedom (SDOF) models to finite element (FEM)
representation. In either case, numerical computation
requires adequate resolution in space and time to account for the high-intensity, short-duration loading
and non-linear response. The challenges are principally
the selection of the model, the anticipated appropriate failure modes, and the interpretation of results for
structural design details. Whenever possible, results
are checked against data from tests and experiments
on similar structures and loadings. Available computer

ARCHITECTURAL PRECAST CONCRETE

|

501

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.5 Blast Effects Predictions

programs include:
• AT Planner (U.S. Army Engineer Research and
Development Center)
• BEEM (Technical Support Working Group)
• BLASTFX (Federal Aviation Administration)
Components such as beams, slabs, or walls can often be modeled by a SDOF system and the governing
equation of motion solved by using numerical methods. There are also charts developed by J. M. Biggs
in Introduction to Structural Dynamics (McGraw-Hill
Publishing Company, 1964) and military handbooks
for linearly decaying loads, which provide the peak response and circumvent the need to solve differential
equations. These charts require only knowledge of the
fundamental period of the element, its ultimate resistance force, the peak pressure applied to the element,
and the equivalent linear decay time to evaluate the
peak displacement response of the system. The design
of the anchorage and supporting structural system can
be evaluated by using the ultimate flexural capacity
obtained from the dynamic analysis. Other charts are
available that provide damage estimates for various
types of construction based on peak pressure and peak
impulse based on analysis or empirical data. Military
design handbooks typically provide this type of design
information.

502

|

creased confidence in the reliability of the results. In
some cases, an SDOF approach will be used for the
preliminary design and a more sophisticated approach,
using finite elements, will be used for the final verification of the design.
A dynamic non-linear approach is more likely than a
static approach to provide a panel cross section that
meets the design constraints of the project. Elastic static
calculations are likely to give overly conservative design
solutions if the peak pressure is considered without the
effect of load duration. By using dynamic calculations
instead of static, it is possible to account for the very
short duration of the loading. Because the peak pressure levels are so high, it is important to account for
the short duration of the loading to properly model the
structural response. In addition, the inertial effect included in dynamic computations greatly improves the
accuracy of the calculated response. This is because by
the time the mass is mobilized, the loading is greatly
diminished, enhancing response. Furthermore, by accepting that damage occurs it is possible to account
for the energy absorbed by ductile systems through
plastic deformation. Finally, because the loading is so
rapid, it is possible to enhance the material strength to
account for high strain-rate effects.

For SDOF systems, material behavior can be modeled
using idealized elastic, perfectly plastic, stress-deformation functions, based on actual structural support
conditions and strain-rate–enhanced material properties. The model properties are selected to provide the
same peak displacement and fundamental period as
the actual structural system in flexure. Furthermore,
the mass and the resistance functions are multiplied by
mass and load factors, which estimate the actual portion of the mass or load participating in the deflection
of the member along its span.

Both concrete and reinforcing steel subjected to the
very short duration impulse type loading caused by a
blast exhibit a higher strength than when subjected
to a static loading. The stiffness and strength of both
steel reinforcement and concrete are likely to increase
with the higher rate of loading under blast conditions.
This obviously increases the strength of reinforced concrete members, which translates into higher dynamic
resistance. But the high rate of loading expected during blasts is also likely to significantly reduce the deformation capacity and the fracture energy of reinforced
concrete. This translates into reduction of ductility of
reinforced concrete in blast loading situations.

For more complex elements, the blast consultant
must resort to finite-element numerical time integration techniques. The time and cost of the analysis must
be considered when choosing design procedures.
SDOF models are suitable for numerical analysis on
PCs, but the most sophisticated FEM systems (with
non-linear material models and options for explicit
modeling of reinforcing bars) may require significant
computing power. Because the design analysis process
is a sequence of iterations, the cost of analysis must
be justified in terms of benefits to the project and in-

In dynamic non-linear analysis, response is evaluated
by comparing the ductility (that is, the peak displacement divided by the elastic limit displacement) and/or
support rotation (the angle between the support and
the point of peak deflection) to empirically established
maximum values that have been established by the
military through explosive testing. Note that these values are typically based on limited testing and are not
well defined within the industry at this time. Maximum
permissible values vary, depending on the material and
the acceptable damage level.

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.5 Blast Effects Predictions

If static design methods are used, it is recommended
that an equivalent static pressure be used rather than
the peak air-blast pressure. The peak air-blast pressure
generally leads to over-designed sections that are not
cost effective, add weight to the structure, and are difficult to construct.
Specifications for precast concrete elements can be
either in the form of a performance requirement, with
the air-blast pressures and required performance provided, or as a prescriptive specification with equivalent
static pressures provided. The equivalent static pressures are computed based on the peak dynamic response of the panel for the defined threat. The performance specifications give precasters more flexibility to
provide the systems with which they are most familiar.
However, it requires that the precaster either have inhouse dynamic analysis capability or have a relationship with a blast engineer who can work with him or
her to customize the most cost-effective system.
On the other hand, as static equivalent pressures are
based on the specific panel’s response to the air-blast
load, changing dimensions, reinforcement, or supported elements would require recalculation of the static
equivalent load and are therefore not recommended
when static equivalent pressures are given as part of
the panel design criteria. However, when using the
static equivalent loads, the designer may proceed normally with the lateral design process, using a load factor of one.
Note that equivalent static values are different from
quasi-static values that assume a displacement ductility
of less than one. The equivalent static values are based
on computations that are non-linear, with ductilities in
excess of one.
Levels of damage computed by means of analysis may
be described by the terms minor, moderate, or major,
depending on the peak ductility, magnitude of support
rotation, and collateral effects. A brief description of
each damage level is given below.
Minor: Nonstructural failure of building elements
such as windows, doors, curtain walls, and false ceilings. Injuries may be expected, and fatalities are possible but unlikely.
Moderate: Structural damage is confined to a localized area and is usually repairable. Structural failure
is limited to secondary structural members such as

5

beams, slabs, and non-loadbearing walls. However,
if the building has been designed for loss of primary
members, localized loss of columns may be accommodated. Injuries and some fatalities are expected.
Major: Loss of primary structural components such
as columns or transfer girders precipitates loss of additional adjacent members that are adjacent to or above
the lost member. In this case, extensive fatalities are
expected. Building is usually not repairable.
Generally, moderate damage at the design threat level is a reasonable design goal for new construction for
which design of blast effects has been specified.
Table 5.6.2 Damage Approximations.

Damage
Typical window glass breakage

Incident
Overpressure
(psi)
0.15 – 0.22

Minor damage to some buildings

0.5 – 1.1

Panels of sheet metal buckled

1.1 – 1.8

Failure of concrete block walls

1.8 – 2.9

Collapse of wood framed
buildings

Over 5.0

Serious damage to steel framed
buildings

4–7

Severe damage to reinforced
concrete structures

6–9

Probable total destruction of
most buildings

10 – 12

SOURCE: Explosive Shocks in Air, Kinney & Grahm, 1985; Facility Damage
and Personnel Injury from Explosive Blast, Montgomery & Ward, 1993; and
The Effects of Nuclear Weapons, 3rd Edition, Glasstone & Dolan, 1977

Figure 5.6.4 provides a quick method for predicting the expected overpressure (expressed in psi) on a
building for a specific explosive weight and standoff
distance. Enter the x-axis with the estimated explosive weight a terrorist might use and the y-axis with
a known standoff distance from a building. By correlating the resultant effects of overpressure with other
data, the degree of damage that the various components of a building might receive can be estimated.
The vehicle icons in Fig. 5.6.4 and 5.6.5 indicate the
relative size of the vehicles that might be used to transport various quantities of explosives.
Figure 5.6.5 shows an example of a range-to-effect
chart that indicates the distance or standoff to which

ARCHITECTURAL PRECAST CONCRETE

|

503

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.5 Blast Effects Predictions/ 5.6.6 Standoff Distance

a given bomb size will produce a given
effect. This type of chart can be used to
display the blast response of a building
component or window at different levels
of protection. It can also be used to consolidate all building response information
to assess needed actions if the threat
weapon-yield changes. For example, an
amount of explosives are stolen and indications are that they may be used against
a specific building. A building-specific
range-to-effect chart will allow quick determination of the needed standoff for
the amount of explosives in question, after the explosive weight is converted to
TNT equivalence.

Fig. 5.6.4 Incident overpressure measured in pounds per square
inch, as a function of stand-off distance and net explosive
weight (pounds-TNT).
Explosives Environment

|

ARCHITECTURAL PRECAST CONCRETE

Trucks
0.5
Incident
Overpressure
(psi)
1.0

1,000
2.0

500
10.0

5.6.6 S tandoff Distance

504

Vans

1,500

0

0

100

1,000

10,000

100,000

Net Explosive Weight (lbs-TNT)

Source: Federal Emergency Management Agency. Reference Manual to Mitigate
Potential Terrorist Attacks. FEMA 426 (Washington, DC: Federal Emergency
Management Agency, December 2003).

Fig. 5.6.5 Explosive environments – blast range to effects.
Explosives Environment

Luggage

10,000

Minimum Stand-off (ft)

Protection for a commercial building, which comes in active and passive
forms, will impact the potential damage
sustained by the building and the rescue
efforts of the emergency workers. The
primary approach is to create a standoff
distance that ensures a minimum guaranteed distance between the blast source
and the target structure. The standoff distance is vital in the design of blast resistant structures since it is the key parameter that determines, for a given bomb
size or charge weight, the blast overpressures that load the building cladding and
its structural elements. The blast pressure
is inversely proportional to the cube of
the distance from the blast to the point in
question. For example, if the standoff distance is doubled the peak blast pressure is
decreased by a factor of eight (Fig. 5.6.6).
Furthermore, for a similar charge weight,
the greater standoff distance results in a
slightly longer loading duration than the
shorter standoff distance, and the blast
wave is more uniformly distributed across
the building face. Currently design criteria
for standoff distances for blast protection
vary from 33 to 148 ft (10 to 45 m) depending on the function of the building.
This standoff distance, or setback zone,
is achieved by placing anti-ram bollards,

Automobiles

2,000

Stand-off Distance (ft)

5

Automobiles

Vans

Trucks

1,000

10
10

100

1,000

10,000

100,000

Weapon Yield (lbs-TNT)
Glass – Minor Cuts

Threshold Injuries – Open or Buildings

Glass – Severe Wounds

Potentially Lethal Injuries

Glass with Fragment Retention
Film – Severe Wounds

Threshold, Concrete Columns Fail
Wall Fragment Injuries or Injuries to
Personnel in Open

KEY CONCERNS ARE GLASS SHARDS AND STRUCTURAL COLLAPSE

Source: Federal Emergency Management Agency. Reference Manual to Mitigate
Potential Terrorist Attacks. FEMA 426 (Washington, DC: Federal Emergency
Management Agency, December 2003).

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.6 Standoff Distance

Fig. 5.6.6 Pressure vs. Range-Hemispherical Surface Burst.*
Pressure vs. Range
Hemispherical Surface Burst
200,000
100,000

Incident Pressure, psi
Reflected Pressure, psi

10,000

Pressure, psi

1,000

100

10

Charge weight
2,000 pounds C-4
Eqv. weight of TNT 2,560 pounds

1

0.2
2

5

10

50

100

500

1,000

2,000

Range, feet

*Bridge and Tunnel Vulnerability Workshop, U.S. Army Engineer Research
and Development Center, Vicksburg, MS, May 13-15, 2003.

5

large planters, low-level walls, fountains, and other barriers that cannot be compromised by vehicular ramming
at the site perimeter (Fig. 5.6.7). The anti-ram capability should be consistent with the size of the weapon
and the maximum achievable velocity of the ramming
vehicle (up to 50 mph [80.5 km/hr]). In urban areas, the
setback choices are limited. In suburban or rural areas,
large setbacks around a building can be used by existing
infrastructure.
The maximum ramming vehicle speeds attainable will
be determined by the site conditions; therefore, the
site conditions will determine the vehicle kinetic energy resulting from an impact that must be resisted
by the standoff barriers. Both the bollard and its slab
connection must be designed to resist the impact loading at the maximum speed attainable. Conversely, if
design restrictions limit the capacity of the bollard or its
slab connection, then site restrictions will be required
to limit the maximum speed attainable by the potential
bomb delivery vehicle.
Fig. 5.6.7 Barriers to achieve standoff distance and limit vehicle
access.
Nimitz-MacArthur Pacific Command Center, Oahu, Hawaii;
Architect: Wimberly Allison Tong & Goo Design;
Photo: Gary Hofheimer Photography.

ARCHITECTURAL PRECAST CONCRETE

|

505

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.6 Standoff Distance/ 5.6.7 Design Concepts

While the setback zone is the most effective and efficient measure to lessen the effect of a terrorist vehicle
bomb attack, it also can work against rescue teams
since the barriers could deter access to the rescue and
firefighting vehicles. In most urban settings, the typical
setback distance from the street to the building façade
is typically 10 to 25 ft (3 to 7.5 m), which does not pose
any access problems for emergency vehicles. However,
when designing prestigious buildings, including landmark office towers, hospitals, and museums, the setback is often increased to 100 ft (30.5 m) or more to
create a grand public space. Details to allow emergency
access should be included in the design of operational
bollards or fences. If plaza or monumental stairs are
used, some secondary access must be incorporated to
similarly allow emergency response entry. Furthermore,
public parking lots abutting the protected building must
be secured or eliminated, and street parking should
not be permitted adjacent to the building. Additional
standoff distance can be gained by removing one lane
of traffic and turning it into an extended sidewalk or
plaza. However, the practical benefit of increasing the
standoff depends on the charge weight. If the charge
weight is small, this measure will significantly reduce
the forces to a more manageable level. If the threat is
a large charge weight, the blast forces may overwhelm
the structure despite the addition of 9 or 10 ft (2.5 to
3 m) to the standoff distance, and the measure may
not significantly improve survivability of the occupants
or the structure.
Figures 5.6.6 and 5.6.8 illustrate the effect of increased standoff distances on the pressures that would
be created on the structure.
Even where the minimum standoff distances are
achieved, many aspects of building layout and other
architectural design issues must be incorporated to improve overall protection of personnel inside buildings.

5.6.7 D
 esign Concepts
Several important concepts should be kept in mind
while designing buildings for blast resistance. These
concepts include energy absorption, safety factors,
limit states, load combinations, resistance functions,
structural performance considerations, and, most importantly, structural redundancy to prevent progressive
collapse of the building. A design satisfying all required
strength and performance criteria would be unsatisfactory without redundancy.

506

|

ARCHITECTURAL PRECAST CONCRETE

Structures with three or more stories are more likely to
be subject to significant damage as a result of progressive collapse. The engineer of record needs to design
the structure to sustain local damage with the structural system as a whole remaining stable and not being
damaged to an extent disproportionate to the original
localized damage. This is achieved through structural
elements that provide stability to the entire structural
system by transferring loads from any locally damaged
region to adjacent regions capable of resisting those
loads without collapse. Transfer girders and the columns supporting them are particularly vulnerable to
blast loading because they support a significant portion of the building above. Unless specially designed,
this form of construction poses a significant impediment to the safe redistribution of the load in the event
the girder or the columns supporting it are damaged.
To limit the extent of collapse of adjacent components: (1) highly redundant structural systems are designed; (2) the structure is analyzed to ensure it can
withstand removal of one primary exterior vertical or
horizontal load-carrying element (a connection, column, beam, or a portion of a loadbearing/shear wall
system) without progressive collapse; (3) connections
are detailed to provide continuity across joints equal to
the full structural capacity of connected members (see
Article 16.5-Structural Integrity in ACI 318); (4) floors
are designed to withstand load reversals due to explosive effects; and (5) exterior walls employ one-way wall
elements spanning vertically to minimize blast loads on
columns.
Strength and ductility (energy-dissipating capacity) are
necessary to achieve high energy absorption, which is
achieved through the use of appropriate structural materials and details. These details must accommodate
relatively large deflections and rotation in order to provide redundancy in the load path. Elements with low
ductility are undesirable for blast resistant design.
Margins of safety against structural failure are
achieved through the use of allowable deformation
criteria. Structures subjected to blast load are typically
allowed to undergo plastic (permanent) deformation
to absorb the explosion energy, whereas response to
conventional loads is normally required to remain in the
elastic range. The more deformation the structure or
member is able to undergo, the more blast energy that
can be absorbed. As member stresses exceed the yield
limit, stress level is not appropriate for judging member
response as is done for static elastic analysis. In dynam-

0
1.6

50

100

150

200

2.8

3.2
3.6

4

4.4

4.8
5.2

100'

Incident

Time, milliseconds

TOA = 2 msec

2.4
5.6

6

150

300

4 msec

175

350

0

25

50

75

100

125

200

250

225

275

42
300

400

Charge weight
2,000 pounds C-4
Eqv. weight of TNT 2,560 pounds
Range
20 feet
Peak pressure
555.1 psi
Impulse
257.8 psi-mse c
Time of arrival
1.915 msec
Duration
3.728 msec
Decay coefficient 0.5436

39

450

2

36

258 psi-msec

0
33

250

555 psi

Hemispherical Surface Burst

5

10

15

20

25

30

35

40

45

50

53 psi

Reflected Pressure

48

51

54

Time, milliseconds

45

57

63

0
66

40

80

120

160

200

240

280

320

360

400

440

Impulse, psi-msec

Reflect ed

60

29 msec

Charge weight
2,000 pounds C-4
Eqv. weight of TNT 2,560 pounds
Range
100 feet
Peak pressure
52.54 psi
Impulse
355.7 psi-msec
Time of arrival
35.09 msec
Duration
28.88 msec
Decay coefficient 10.1

356 psi-msec

Hemispherical Surface Burst

500

550

600

Incident Pressure

Pressure, psi

*Bridge and Tunnel Vulnerability Workshop, U.S. Army Engineer Research and Development Center, Vicksburg, MS, May 13-15, 2003.

20'

Ground Level Bomb

2,000 lb C-4

Pressure, psi

55

Impulse, psi-msec

Fig. 5.6.8 Explosive Airblast Loadings from Vehicle Bombs.*

0
33

2

4

6

39

42

45

48

51

54

TOA = 35 msec

Incident

Time, milliseconds

60

63

0
66

15

30

45

Rear Face

57

29 msec

75

10

60

90

12

8

105

135

14

Charge weight
2,000 pounds C-4
Eqv. weight of TNT 2,560 pounds
Range
100 feet
Peak pressure
18.03 psi
Impulse
145.7 psi-msec
Time of arrival
35.09 msec
Duration
28.88 msec
Decay coefficient 14.02

146 psi-msec

150

120

36

18 psi

Incident Pressure
Hemispherical Surface Burst

16

18

20

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.7 Design Concepts

ARCHITECTURAL PRECAST CONCRETE

|

5

507

Impulse, psi-msec

Pressure, psi

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.7 Design Concepts

ic design, the adequacy of the structure is judged on
maximum deformation capability. Limits on displacements are selected based on test data or other empirical evidence as well as blast probability and potential
consequences. A degree of conservatism is included to
ensure adequate capacity because the applied loads are
not “factored up” to provide a factor of safety.
As long as the calculated deformations do not exceed
the allowable values, a margin of safety against failure
exists. Since the actual weight of the explosive charge
is threat/risk based, the engineer cannot increase the
design blast pressure loading and be assured of achieving a margin of safety. Blast-resistant design requires
that the loads from blasts be quantified by risk analysis
to determine the threat (charge weight) and that the
structural performance requirements be established for
buildings subjected to these derived loads. Methods to
determine the blast loading and structural performance
limits are established in TM 5-1300 for buildings exposed to explosions from TNT or other high-yield explosives in military applications and munitions plants.
Typical threats for civilian structures vary from suitcase
and backpack bombs (20 to 50 lbs [9 to 23 kg] TNT
equivalent) to van or small truck bombs (3000 to 5000
lbs [1360 to 2267 kg] TNT equivalent). Generally, the
larger charge sizes are associated with vehicles that can
be kept farther from the building (60 to 100 ft [18 to
30.5 m]) by appropriately designed vehicle barriers.
Design codes contain special provisions for high seismic conditions, which may be used to address the requirements to design against progressive collapse associated with design for blast resistance. However, these
provisions are not sufficient for blast design. These provisions are intended to protect against nonductile failure modes, such as buckling or premature crushing of
brittle materials, through use of special detailing and
design requirements. The desirable features of earthquake-resistant design (ductility, redundancy, and load
redistribution) are equally desirable in blast design. The
provision for seismic detailing, which maintains the
capacity of the section despite development of plastic
hinges, is also desirable for resisting the effects of blast.
However, the highly localized loading from a blast and
the potential for different mechanisms/failure modes
requires some additional considerations. For blast effects, the engineer should design the panels so that
the full capacity of the section will be realized and that
no premature failure will occur.

508

|

ARCHITECTURAL PRECAST CONCRETE

Building codes define the load factors and combinations of loads to be used for conventional loading
conditions such as dead, live, wind, and earthquake.
However, no current building codes cover blast loading
conditions. Blast loads are combined with only those
loads that are expected to be present at the time of
the explosion. Therefore, blast loads are not combined
with earthquake or wind loads.
The Strength Design Method of ACI 318 may be
used to extend standard concrete strength and ductility requirements to the design of blast resistant structures. The resistance of concrete elements subjected to
high strain rates is computed using dynamic material
strengths, which are typically 10 to 30% greater than
static load strengths. Strength reduction or resistance
factors are not applied (ø = 1.0) to load cases involving
blast. The plastic response used in blast design is similar
in concept to the moment redistribution provisions in
ACI 318, Section 8.4 and the seismic criteria provided
in ACI 318, Chapter 21. The seismic detailing provisions
are applied to provide the necessary ductile response.
In addition to ACI 318 requirements, the following
items should be considered for blast resistant design:
1. The minimum reinforcing provisions of ACI 318
apply; however, the option to use one-third more
reinforcing than computed should not be taken.
The moment capacity of under-reinforced concrete
members is controlled by the uncracked strength
of the member. To prevent a premature ductile failure, reinforcing in excess of the cracking moment
should be provided. Two-way, symmetric reinforcement is recommended to accommodate large deformations and rebound loads.
For panels, the minimum reinforcement ratio (reinforcing steel cross-sectional area to the panel crosssectional area) of vertical reinforcing steel should
be equal to or greater than ACI 318 minimums required for Seismic Design Categories D, E, or F. If
the risk potential for a blast is high, the minimum
reinforcement ratio required for blast-resistant design (TM 5-855-1; DAHSCWEMAN 1998) should
be used as a basis for design. Generally, for concrete walls 8 in. (200 mm) or greater in thickness,
the recommended minimum reinforcing should
be 0.25% each face. For concrete walls less than
8 in. (200 mm) thick, 0.5% as a single layer (on
center line) of reinforcing should be the minimum
specified.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.7 Design Concepts / 5.6.9 Designing Precast Concrete Panels

2. Code provisions for maximum allowable reinforcing are included to prevent crushing of concrete
prior to yielding of steel. Code provisions also allow compression reinforcing to offset maximum
tension reinforcing requirements. Because blast
resistant precast concrete panels typically have the
same reinforcing on each face to resist rebound
loads, maximum reinforcing provisions should not
be a problem.
3. The substitution of higher grades of reinforcing
should not be allowed. Grade 60 reinforcing bars
(No. 11 [36] and smaller) have sufficient ductility
for dynamic loading. Bars with high yield strength
may not have the necessary ductility for flexural resistance and shop bending; straight bars should be
used when possible for these materials. Welding
of reinforcement is generally discouraged for blast
design applications; however, in some instances
it may be required for anchorage. In these cases,
ASTM A706 bars may be used.
4. Development lengths should not be reduced for
excessive reinforcement. Because plastic hinges
will cause over-designed reinforcing to yield, the
full actual strength of reinforcing should be used in
computing section capacities. The development of
reinforcing should be computed accordingly.
5. Criteria intended to reduce cracking at service load
levels need not be applied to load combinations
including blast. Cracking and permanent deformations resulting from a plastic range response are
expected results of such an unusual type of load.
6. Some concrete elements are simultaneously subjected to out-of-plane bending loads in combinations with in-plane shear loads. For example, side
walls must resist side overpressures acting into the
plane of the side wall. Additionally, reactions from
the roof diaphragm acting in the plane of the side
shear wall must also be resisted.

5

For a surface blast, the most directly affected building elements are the façade and structural members
on the lower four stories. Although the walls can be
designed to protect occupants, a very large vehicle
bomb at small standoffs will likely breach any reasonably sized wall at the lower levels. There is a decrease
in reflected pressure with height due to the increase
in distance and angle of incidence of the air blast at
the upper levels of a highrise building. Chunks of concrete spalled from panels by blast forces move at high
speeds and are capable of causing injuries. Additional
protection from fragment impact can be provided by
steel backing plates, carbon fiber materials, or KEVLAR
lining the interior of the wall; however, these are extreme measures that should be reserved for localized
protection of high-value assets.
The building structure, architectural precast concrete cladding, and the window, window wall, and
any curtain wall framing systems may be designed
to adhere to the blast criteria within the Interagency
Security Committee (ISC) “Security Design Criteria for
New Federal Office Buildings and Major Renovation
Projects,” dated May 28, 2001, for the appropriate
hazard level as determined by a threat consultant. By
combining the criteria of the ISC with the applicable
blast analysis standards mentioned earlier, the architectural precast concrete cladding systems should be
sufficiently sized, reinforced, detailed, and installed to
resist the required blast loading criteria on the panels if they were tested in accordance with the General
Services Administration’s (GSA) “Standard Test Method
for Glazing and Window Systems Subject to Dynamic
Overpressure Loadings” (GSA – TS01-2003). In addition to safely transferring the blast pressures into the
supporting structure, the panels also must be checked
for their capacity to transfer the additional loading
caused by the specified window framing and the blastresistant glass units.

5.6.8 F açade Considerations

5.6.9 Designing Precast Concrete Panels

A major structural consideration is the construction
of the exterior façade. Second only to the impact the
standoff distance has on the effects of the blast, the
façade remains the occupants’ last form of true protection. Not only does the building’s skin protect the occupants from the weather, but it also has the potential
to limit the blast pressure that can actually enter the
workspace.

Architectural precast concrete can be designed to
mitigate the effects of a bomb blast and thereby satisfy GSA and DOD requirements. Rigid façades, such
as precast concrete, provide needed strength to the
building through in-plane shear strength and arching action. However, this strength usually is not taken
into consideration in conventional design, as design
requirements do not require those strength measures.

ARCHITECTURAL PRECAST CONCRETE

|

509

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.9 Designing Precast Concrete Panels

Panels are designed for dynamic blast loading rather
than the static loading that is more typical. Precast
concrete walls, being relatively thin flexural elements,
should be designed for a ductile response (eliminating
brittle modes of failure). There are tradeoffs in panel
stiffness and the forces that must be reacted to by the
panel connections that must be evaluated by the engineer. Typically, the panels should have increased section
thickness or ribs on their back side and have as much
as 75% additional reinforcement. But, the amount of
flexural reinforcing should be limited to ensure that
tensile reinforcing will yield before concrete crushing
can occur. Shear steel may be used to increase shear
resistance, confine the flexural reinforcing, and prevent
buckling of bars in compression.
For precast concrete panels, designers should consider a minimum thickness of 5 in. (125 mm) exclu-

sive of reveals, with two-way reinforcing bars spaced
not greater than the thickness of the panel. For thin
panels, where it is difficult to place two layers of reinforcement, the use of two layers of heavy welded wire
reinforcement along the centerline or staggered bars
on either face may be considered. These reinforcement
details will improve ductility and reduce the chance of
flying concrete fragments. The objective is to reduce
the loads transmitted into the connections, which
need to be designed to resist the loads associated with
the ultimate flexural capacity of the panels.
Precast concrete panels are subject to horizontal loadings due to wind, earthquake, and blast and in-plane
loads due to earthquakes. As a means of addressing
these loads, they may be analyzed separately. This is a
satisfactory design approach based on the International
Building Code (IBC) load combinations.
Deep surface profiling of panels should be minimized;
such features can enhance blast effects by causing complex reflections and lead to a greater level of damage
than would be produced with a non-profiled façade.
To accommodate blast loading, the following features are commonly incorporated into precast concrete
panel systems:
1. Increase panel size to at least two stories tall or
one bay wide to increase the ductility. Panels can
then absorb a larger portion of the blast energy
and transfer less through connections to the main
structure. Typically, the largest panel is analyzed
for wind, seismic, and dead-loading and connections for all the panels are based on those results.
But with bomb-blast criteria, the goal is to provide
panels with the flexibility to bend, break, or crush
while remaining essentially intact. As a result, often
the smaller, less flexible panels in each group may
be the critical components, and these are analyzed
for loading instead.

Fig. 5.6.9 The 6 in. thick x 22 ft.
tall panels were reinforced with
ribs spaced 6 ft. apart.

510

|

ARCHITECTURAL PRECAST CONCRETE

2. Panels should be connected to floor diaphragms,
rather than to columns, in order to prevent overstressing of the columns. The panels would then
fail individually. When panels are connected to the
floors rather than the columns, movement of any
panel during erection causes the previously set and
tied-back panel to lose alignment. The amount of
deflection of the floor or beam varies with the panel’s position on the floor or beam, requiring field estimates to determine how high to set each panel to
allow for deflection caused by the adjacent panel.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.9 Designing Precast Concrete Panels

5

Fig. 5.6.10
Lloyd D. George United States Courthouse, Las Vegas, Nevada; Architect: Dworsky Associates, Design Architect; Langdon Wilson
Architects, Executive Architect; Photo: Langdon Wilson Architects.

3. Panels may be designed with integrally cast and
reinforced vertical pilasters or ribs on the back to
provide additional support and act as beams that
span floor-to-floor to take loads (Fig. 5.6.9). This
rib system makes the panels more ductile and better able to withstand an external blast. But it also
forces the window fenestration into a “punched”
opening symmetry. In addition, greater edge of
slab clearances must be provided to accommodate
the ribs.
Loadbearing precast concrete panels need to be designed to span over assumed failed areas by means of
arching action, strengthened gravity connections, secondary support systems, or other means of providing
an alternate load path. The precast concrete structure
must satisfy the requirements of ACI 318, Sections
7.13.3 and 16.5.

For loadbearing wall structures, the following detailing recommendations on connections/ties will help resist progressive collapse:
• Provide horizontal and vertical ties in vertical joints
between adjacent or intersecting bearing walls.
• Connect panels across horizontal joints by a minimum of two connections per panel.
• Connect all members to the lateral force resisting
system and their supporting members. Tension ties
must be provided in the transverse, longitudinal,
and vertical directions and around the perimeter of
the structure.
• Provide ties between transverse bearing walls and
connecting floor panels.
• Do not use connection details that rely solely on
friction caused by gravity loads.

ARCHITECTURAL PRECAST CONCRETE

|

511

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.9 Designing Precast Concrete Panels

Fig. 5.6.11
United States Federal Courthouse, Jacksonville, Florida;
Architect: KBJ Architects Inc.; Photo: Aerial & Architectural Photo, Inc.

512

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.9 Designing Precast Concrete Panels / 5.6.10 Examples of Projects Designed for Blast

In facilities designed for blast, architectural precast
concrete column covers should be at least 6 in. (150
mm) from the structural member. This will make it
considerably more difficult to place an explosive device
directly against the structure. Because explosive pressures decay so rapidly, every inch of distance will help
to protect the column.

5

9 in. (225 mm) thick with about 15% of the panels 7
in. (175 mm) thick. The panels transfer energy to the
structure at the floor slabs, which act as diaphragms.
This avoids having the blast loads transferred to the
center of the structural columns, which could overstress them. The wind load requirements varied from
63 to 104 psf (3 to 5 kPa) depending on location and
other variables while the dynamic blast criteria varied
from an equivalent static pressure of 288 to 576 psf
(14 to 28 kPa). To resist the increased forces generated
by the blast energy, larger connection hardware than
usual was used.

5.6.10 E xamples of Projects Designed
for Blast
The first major federal courthouse built after the 1995
blast in Oklahoma City contains features intended to
avoid a catastrophic collapse in the event of a terrorist
attack (Fig. 5.6.10). The precast concrete panels were
designed to be more ductile than conventional panels
so they could absorb as much of a bomb blast as possible without destroying the connections that tie them
to the main structure. The 6 in.-thick (150 mm) panels
span 22 ft (6.7 m) from floor to floor. The vertical reinforcing rib system is shown in Fig. 5.6.9. Also note the
use of bollards to increase the standoff distance.

Beyond the typical performance requirements of designing for wind and seismic loads, the precast concrete panels and connections on the speculative office
building in Fig. 5.6.12 had to resist directly applied
blast loads plus the blast loads collected by the adjacent window systems that were superimposed on the

(b)

The 15-story courthouse (Fig. 5.6.11) houses 17
courtrooms for a variety of federal jurisdictions. The
precast concrete panels span floor-to-floor on the
courtroom levels, requiring panels between 17 and 22
ft (5.2 and 6.7 m) tall. The majority of the panels are

(a)

Fig. 5.6.12 (a) & (b)
Patriot Square Washington, D.C.; Architect: Gensler D.C.;
Photos: Maxwell Mackenzie.

ARCHITECTURAL PRECAST CONCRETE

|

513

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.10 Examples of Projects Designed for Blast

(b)

(a)

Fig. 5.6.13 (a) & (b)
Fresno Courthouse and Federal Building, Fresno, California; Architect: Moore Ruble Yudell Architects & Planners,
Design Architect; Gruen Associates, Executive Architect.

panels. The building used texture and three-dimensional architectural precast concrete panels to create a
distinctive look. The project uses two colors of precast
concrete, replicating a limestone finish and a white,
contrasting shade, to help create layers on the building

514

|

ARCHITECTURAL PRECAST CONCRETE

and make the plaza areas (in white) stand out. The result makes the building seem even crisper in color and
design. The textured approach was emphasized by the
use of thin strips of precast concrete beneath the ribbon window, adding additional shadow lines.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.10 Examples of Projects Designed for Blast

The courthouse and federal building with nine levels
above ground is the tallest structure in Fresno, Calif.
(Fig. 5.6.13[a]). The patterning of the architectural precast concrete panels was designed to give a human
scale to the potentially imposing building mass. The
combination of vertical and horizontal patterning provides variation in visual texture throughout the day.
The lower floors have a more rugged appearance than
the upper floors and a warmer color to differentiate
the base from the body of the building (Fig. 5.6.13[b]).
The 6-in.-thick (150 mm) precast concrete panels are
reinforced with integrally cast, large-ribbed pilasters to
strengthen the panels and help achieve the blast resistance objectives.

5

(b)

The federal building, totaling 2 million gross sq. ft
(185,800 m2), consists of two separate buildings that
are architecturally similar, eight and nine stories in
height, respectively (Figs. 5.6.14[a] and [b]). While the
complex, with its bold colors of white, dark green, and

(a)
Fig. 5.6.14(a) & (b)
U.S. Department of Transportation Headquarters, Washington, D.C.;
Architect: Michael Graves & Associates, Design Architect; DMJM/AECOM,
Architect of Record; Photos: ©Eric TaylorPhoto.com.

ARCHITECTURAL PRECAST CONCRETE

|

515

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.10 Examples of Projects Designed for Blast

red, provides an outwardly open appearance and perception to the public, it conforms to stringent security criteria, including the GSA security design for new
buildings as well as force protection guidelines (façade
resistance to blast pressures). The building was originally designed to be 20 ft. (6.1 m) from the defensible
perimeter inside the street curb line to the face of the
building; this dimension was increased to 50 ft (15.2
m). The requirement to provide “controlled” vehicular
access to the street between the two buildings was
also added. The design of all of the exterior façade elements provides for the higher peak pressures and impulses for all panels regardless of the increased standoff distance from the explosive device or the calculated
height above the explosive device. Additionally, in
those areas where the standoff distance is less than 50
ft (15.2 m), the design provides for an “equivalent level

of protection” to those exterior façade elements that
would be affected. To provide the equivalent level of
protection, the double-pane insulated glazing with an
interior lite of laminated glazing, adhered to enhanced
mullions, was designed to satisfy the maximum design
peak pressure and impulse loading on the exterior façade elements with the limitation that fragments from
no more than 10% of the total glazed area may exceed the GSA performance criteria.
Design criteria for the walls of the railroad command
center (Fig. 5.6.15) includes the ability to withstand the
impact of an irregular object at 200 mph (322 km/hr)
and equivalent explosive force. The use of 12-in.-thick
(300 mm) precast concrete wall panels met this criteria
and, in their capacity as loadbearing walls, lessened
the building’s cost.

Fig. 5.6.15
CSX (formerly Conrail) Computer Technology Center,
Philadelphia, Pennsylvania;
Architect: KlingStubbins;
Photo: CG Berken.

516

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.11 Connection Concepts and Details

5.6.11 Connection Concepts and Details
Architectural precast concrete construction relies on
mechanical connectors at discrete locations that may
be damaged in a blast event, posing specific design
issues for the engineer. These concerns can be overcome with proper detailing. The governing connection
forces are based on the maximum percentage of reinforcement for wind, seismic, or blast loading, since the
amount of steel is generally proportional to panel stiffness. The reaction forces for the design of the anchorages and connections should be based on panel width
and be considered factored loads. The wind load reactions are based on elastic deformations of the panels.
Precast concrete cladding wall panel connection details
for blast effects may be strengthened versions of conventional connections with a likely significant increase
in connection hardware (Fig. 5.6.16 through 5.6.19).

5

Connection details also may emulate cast-in-place concrete to provide a building that provides element continuity. For a panel to absorb blast energy (and provide
ductility) while being structurally efficient, it must develop its full plastic flexural capacity, which assumes the
development of a collapse mechanism. The connection
failure mode should be yielding of the steel and not
splitting, spalling, or pulling out of the concrete. Design
material strengths may be increased by a dynamic-increase factor because of strain rate enhanced material
properties. For structural steel, the factor ranges from
1.05 to 1.10 for tensile and yield strength depending
on the steel grade. A factor of 1.25 applies to concrete compressive strength and the tensile and yield
strengths of reinforcing steel. Also, the shear capacity
of the component should be at least 20% greater than
the member’s dynamic flexural strength. Steel-to-steel
connections should be designed so the weld is never

Fig. 5.6.16 – 19 Connection details.
Fig. 5.6.16(a – c) Panel to panel or alignment connections.

Upper Panel
Exterior Face

(a)

(b)
Typical Section

Rear View

Lower Panel

(No Shims)

Slotted Insert
(c)

ARCHITECTURAL PRECAST CONCRETE

|

517

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.11 Connection Concepts and Details

Fig. 5.6.17(a – e) Bearing connections.

(b)

(a)

(c)

Clips

Plan

(d)
Section

CL Bolt

Temporary
Safety
Connection

(e)

Section

518

|

ARCHITECTURAL PRECAST CONCRETE

Note the placement of the shim stack
on back side of column pilaster and inline with the welded clip angles. This
accommodates the out of plane rotation
of the panel connection when the panel
bows inward and outward from the blast
pressures.

The welded clip angles are designed to
plastically deform when subjected to
the loads and subsequent rotation of the
panel in response to blast pressures. The
deformation of these angles help to dissipate the energy from the blast.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.11 Connection Concepts and Details

5

Fig. 5.6.18(a) & (b) Push/pull or tie-back connections.

(b)
The plate is free to move through the
vertical slot in the down rigger tube. This
is how story drift is accommodated.

(a)

Plan

These plates are designed to act like
springs to help absorb energy from the
blast loading. They also will deform
plastically which helps dissipate energy.
Section

(c)

Fig. 5.6.19 Column cover connection.

CL Bolt

CL Bolt

Alternating
Joints

Section

Rear View

Adjacent Wall Panel
Plan

ARCHITECTURAL PRECAST CONCRETE

|

519

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.11 Connection Concepts and Details / 5.6.12 Glazing

the weak link in the connection. Coordination with interior finishes needs to be considered due to the larger
connection hardware required to resist the increased
forces generated from the blast energy.
Where possible, connection details should provide for
redundant load paths, since connections designed for
blast may be stressed to near their ultimate capacity, the
possibility of single connection failures must be considered. Consideration should be given to the number of
components in the load path and the consequences of
a failure of any one of them. The key concept in the development of these details is to trace the load or reaction through the connection. This is much more critical
in blast design than in conventionally loaded structures.
Connections to the structure should have as direct a
load transmission path as practical, using as few connecting pieces as possible.
Rebound forces (load reversal) can be very high. These
forces are a function of the mass and stiffness of the
member as well as the ratio of blast load to peak resistance. A connection that provides adequate support
during a positive phase load could allow a member to
become dislodged during rebound. Therefore, connections should be checked for rebound loads (even if the
panel is not designed for rebound). It is conservative to
use the same load in rebound as for the inward pressure. More accurate values may be obtained through
dynamic analysis and military handbooks.
It is also important that connections for blast-loaded
members have sufficient rotational capacity. A connection may have sufficient strength to resist the applied
load; however, when significant deformation of the
member occurs this capacity may be reduced due to
buckling of stiffeners, flanges, or changes in nominal
connection geometry, etc.
The capacity of a panel to deform significantly and
absorb energy is dependent on the ability of its connections to maintain integrity throughout the blast response. Failure can occur if connections become unstable at large displacements. The overall resistance of the
panel assembly will be reduced, increasing deflections
or otherwise impairing panel performance.
Both bolted and welded connections can perform
well in a blast environment if they can develop strength
at least equal to that of the connected elements (or at
least the weakest of the connected elements).

520

|

ARCHITECTURAL PRECAST CONCRETE

5.6.12 Glazing
The façade is comprised of the transparent glazing
and opaque exterior wall elements. The glazing, a blastsensitive element, is the first building component likely
to fail in response to the initial blast pressure that engulfs the building. Although the opaque wall elements
may be designed to resist the loading, the options available for the glass are much more limited. These options
include selecting an appropriate type of glass, applying security window (fragment-retention) film, installing blast curtains/shields, and/or using laminated glass.
Due to the extreme intensity of car bomb blast pressures, glazing on the blast side of the target structure
will likely fail depending on the standoff distance. There
is a direct correlation between the degree of fenestration and the amount of debris that enters the occupied
space. Historically, failed window glazing due to the direct pressures produced by an explosion has resulted
in a considerable proportion of injuries, casualties, and
loss of use of the facility.
The two keys to protecting the workspace are attempting to prevent the windows from failing and
ensuring that the windows fail properly if overloaded.
While a great number of injuries are related to flying
glass shards, it is not the only significant source of injury, though it is usually a more visible one. The other
visible cause of injury is falling debris. One of the less
visible causes of injuries is blast pressure, which can
rupture ear drums, collapse lungs, or even crush skulls.
These injuries, which begin at pressures near 15 psi
(103 kPa), can be reduced if the level of blast pressures
entering the space is reduced. The amount of blast
pressure that enters the space is directly proportional
to the amount of openings on the structure’s façade.
Also, smaller windows will generally break at higher
pressures than larger windows, making them less
prone to breakage. Consideration should be given to
designing narrow, recessed windows with sloped sills
Fig. 5.6.20 Narrow and recessed windows with sloped sills.

Source: U.S. Air Force, Installation Force Protection Guide

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.12 Glazing

because they are less vulnerable to blast (Fig. 5.6.20).
Narrow, recessed windows, however, will impact the
building’s design both aesthetically and functionally. To
the extent that nonfrangible glass isolates a building’s
interior from blast shock waves, it can also reduce
damage to interior framing elements (supported floor
slabs could be made to be less likely to fail due to uplift
forces) for exterior blasts.
In embassies, the earliest type of civilian building designed to resist blast events, fenestration is limited to
15% of the effective wall area (calculated using the
floor-to-floor height and width of a single bay). While
this helps in the protective design, it does not provide
the proper lighting or open feeling that is desired in
commercial office buildings; consequently, the fenestration limitations are often increased to 40% for commercial buildings with increased scrutiny paid to glazing detailing.
The second design aspect for windows is to ensure
that they fail properly if overloaded. Special blast-resistant windows can be designed not to fail for the
small to mid-sized opening described previously, provided that the loading is limited. Annealed float glass,
the most common form of architectural glass, behaves
poorly when loaded dynamically.
While typical annealed float glass is only capable of
resisting, at most, 2 psi (14 kPa) of blast pressure, several other types of glazing exist that can resist moderately larger blast pressures. Thermally tempered glass
(TTG) (ANSI Z97.1 or ASTM C1048) and polycarbonate
glazing (also known as bullet-resistant glass) can be
made in sheets up to about 1 in. (25 mm) thick and can
resist pressures up to about 30 to 40 psi (200 to 275
kPa). Laminated (60 mil interlayer thickness) annealed
glass with a 1.4 in. (6 mm) bead of structural sealant
around the inside perimeter exhibits the best postdamage behavior and provides the highest degree of
safety to occupants. (Refer to ASTM F2248, Standard
Practice for Specifying an Equivalent 3-Second Duration
Design Load for Blast Resistant Glazing Fabricated
with Laminated Glass). The lamination holds the glass
shards together in explosive events, reducing their
potential to cause laceration injuries. The structural
sealant helps to hold the pane in the frame for higher
loads. For insulated units, only the inner pane needs to
be laminated. Associated with each of these upgrades
is a considerable increase in cost for the glazing material. The window bite (the depth of window captured
by the frame) needs to be at least 1/2 in. (13 mm).

5

Equally important to the design of the glass is the
design of the glazing system and the framing to which
the glazing is attached. Glazing, frames, and attachments must be treated as an integrated system and be
capable of resisting blast pressures and transferring the
loads to the cladding to which the frame is attached.
To fail as predicted, a window must be held in place
long enough to develop the proper stresses that cause
failure. Otherwise, the window may disengage from
its frame intact and pose a post-event threat or cause
serious damage or injury. Therefore, the frame and anchorage should be designed to develop the full loading
anticipated for the chosen glazing type. Depending on
the façade, the cladding panels to which the windows
are attached must be able to support the reaction forces of a window loaded to failure.
Window frames and mullions of steel, steel-reinforced
aluminum, and heavy-walled aluminum are common
for blast-resistant framing components. Frames, mullions, and window hardware should be designed to
resist a minimum static load of 1 psi (7 kPa) applied
to the surface of the glazing or a dynamic load may
be applied using the peak pressure and impulse values. However, designing for 1 psi static loading will
not necessarily ensure that the window frames, mullions, and anchorages are capable of developing the
full strength of the laminate interlayer. The equivalent
static value is dependent on the type of glass, thickness of glass, size of window unit, and thickness of
laminate interlayer used. Also, a static approach may
lead to a design that is not practical, as the mullion can
become very deep and heavy, driving up the weight
and cost of the window system.
The loading of the frame will depend on the design
blast pressure and the size of the window. As a minimum, frame connections to surrounding walls should
be designed to resist a combined ultimate loading
consisting of a tension force of 200 lbs/in. (35 kN/m)
and a shear force of 75 lbs/in. (13 kN/m). Typically,
this requires a plate with anchors rather than a simple
bolted connection. Frame-supporting elements and
their connections should be designed based on their
ultimate capacities. In addition, because the resulting dynamic loads are likely to be dissipated through
multiple mechanisms, it is not necessary to account for
reactions from the supporting elements in the design
of the remainder of the structure. Additional reinforcement should be provided at window openings. Vertical
and horizontal reinforcement that would have occu-

ARCHITECTURAL PRECAST CONCRETE

|

521

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.12 Glazing

pied the opening width should be evenly distributed
on each side. Also, shear reinforcement should be provided as required around the opening.

The outer frame, referred to as an embed, is fabricated from A36 plate, channel, or angle depending on
the particular geometry of the concrete wall and architectural treatment. The embed shown in Fig. 5.6.21
consists of a 1/2 in. × 6 in. (13mm × 150mm) steel plate.
The inner frame is connected to the embed using highstrength bolts in drilled and tapped holes in the embed
plate. Shim space should not be greater than 1/4 in.
(6.3mm) to minimize the length of the frame bolts.
Corrosion resistant, usually stainless, shims are placed
at each bolt when required. The frames may be cantilevered out from the edge of the wall to reduce the
recessed distance when a thick architectural façade is
used. This cantilevered distance is usually not greater
than 1.5 in. (38mm).

Figure 5.6.21 shows a typical section through a frame
containing a blast window. The primary elements include an inner frame holding the glazing and an outer
frame anchored to the structure. The inner frame consists of a frame angle and glazing stop. The frame angle is typically an A36 angle cut to the desired dimensions. The glazing stop is fabricated from a structural
angle, a structural tube (as shown), or an A36 bar with
countersunk holes. The entire inner frame is designed
to allow replacement of the glazing. Windows are typically factory-glazed and mounted in the window openings as a complete unit.

The blast-resistant glazing for the Lloyd D. George
Federal Building and United States Courthouse, Las
Vegas is a 1 in. (25 mm) thick insulating unit composed
of an annealed exterior light, a 1/2 in. (13 mm) air space,
and a laminated interior lite held in place by an aluminum frame (Fig. 5.6.22. The inboard lite is composed
of a polyvinyl-butral layer between two sheets of 1/8 in.
(3 mm) thick annealed glass. This design uses annealed

The window is held and supported by continuous
gaskets on the inside and outside faces of the glazing.
Neoprene gaskets are used for glass and santoprene
is used for polycarbonate/glass lay-ups. Setting blocks
provide a cushion for the glazing and clearance for
thermal expansion and rotation of the glazing during
blast loading.
Fig. 5.6.21 Generic blast window glazing and frame detail.
INTERIOR

EXTERIOR

Tube
Glazing
Stop

Glazing

Frame Bolts

Frame Angle

Embed Plate
Shims
RC Wall

Shear Stud

Source: Structural Design for Physical Security: State of the Practice, Structural Engineering Institute of American Society of Civil Engineers, Reston, VA, 1999.

Fig. 5.6.22 Blast-resistant glazing detail.
1 Concrete panel with embedded steel channel
2 Screw fastener
3 Gasket with silicone sealant above
4 Sacrificial lite
5 Blast-resistant glass
6 Steel plate
7 Bolt

522

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.12 Glazing / 5.6.13 Initial Costs

glass in lieu of the stronger tempered glass because it
has more flexible properties, which absorb the impact
of the explosion.
Window glazing assessments and designs for blast
response may be performed using one of the government produced and sponsored computer programs such as WINGARD (WINdow Glazing Analysis
Response & Design). This computer program was developed by the US General Services Administration
and is available to Government Agencies and their
contractors. WINGARD may be downloaded from the
GSA’s Office of the Chief Architect web site (www.
oca.gsa.gov) or obtained from the developer (Applied
Research Associates, 119 Monument Place, Vicksburg,
MS 39180). The engineer should define the structural
design criteria and coordinate with the building’s architect to assure the window manufacturer’s correct
interpretation.
Drawbacks of high-performance glazing systems include cost and high maintenance. When the cost for
installing blast-resistant windows is significant relative
to the total cost of the building, resources allocated
to protective design may be better applied toward
upgrading the structural frame to be blast resistant.
This is because the blast pressures from a close in car
or truck bomb can far exceed the allowable pressures
any window system can resist. As a point of reference,
façade blast pressures in the Oklahoma City bombing
were on the order of 4,000 psi (28 MPa) – 100 times
higher than the design pressures described above.
Atriums incorporating large vertical glazed openings
on the building façade, common in prestigious office
buildings, cannot be designed to withstand blast pressures from a close-in explosion. It is not reasonable to
harden the exterior walls of the structure and leave an
atrium’s exterior wall of this type as an inviting target.
Atrium balcony parapets, spandrel beams, and exposed
slabs must be strengthened to withstand loads that are
transmitted through exterior glass or framing. Another
approach is to use an internal atrium with no outward
facing windows or an atrium with clerestory windows
that are close to the ceiling and angling the windows
away from the curb to reduce the pressure levels.

5.6.13 I nitial Costs
The initial construction cost of protection has two
components: fixed and variable. Fixed costs include
such items as security hardware and space require-

5

ments. These costs do not depend on the level of an
attack; that is, it costs the same to keep a truck away
from a building whether the truck contains 500 or
5000 lb (227 or 2268 kg) of TNT. Blast protection, on
the other hand, is a variable cost. It depends on the
threat level, which is a function of the explosive charge
weight and the standoff distance.
The optimal standoff distance is determined by defining the total cost of protection as the sum of the
cost of protection (construction cost) and the cost of
standoff (land cost). These two costs are considered as
a function of the standoff for a given explosive charge
weight. The cost of protection is assumed to be proportional to the peak pressure at the building envelope, and the cost of land is a function of the square of
the standoff distance. The optimal standoff is the one
that minimizes the sum of these costs.
If additional land is not available to move the secured
perimeter farther from the building, the required floor
area of the building can be distributed among additional floors. As the number of floors is increased, the
footprint decreases, providing an increased standoff
distance. Balancing the increasing cost of the structure
(due to the added floors) and the corresponding decrease in protection cost (due to added standoff), it is
possible to find the optimal number of floors to minimize the cost of protection.
Though it is difficult to assign costs to various upgrade measures because they vary based on the site
specific design, some generalizations can be made
(Fig. 5.6.23). In some cases, the owner may decide to
prioritize enhancements, based on their effectiveness
in saving lives and reducing injuries. For instance, measures against progressive collapse are perhaps the most
effective actions that can be implemented to save lives
and should be considered above any other upgrades.
Laminated glass is perhaps the single most effective
measure to reduce extensive non-fatal injuries.
An awareness of a blast threat from the beginning of
a project helps to decide early what the priorities are
for the facility. Including protective measures as part of
the discussion regarding trade-offs early in the design
process often helps to clarify the issues.
Ultimately, the willingness to pay the additional cost
for protection against blast hazards is a function of the
“probability of regrets” in the event a sizable incident
occurs. In some situations, the small probability of an
incident may not be compelling enough to institute

ARCHITECTURAL PRECAST CONCRETE

|

523

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.6.13 Initial Costs / 5.6.14 References

Fig. 5.5.23 Plots showing relationship between cost of upgrading various building compnents, standoff distance, and risk.

Incremental Cost of Protection ($)

Total Protection Cost
(hardening + land + perimeter)
Not to Scale
cost of land
+ perimeter
protection
frame

cost of
hardening

windows and walls

progressive
collapse
other, mailroom, loading dock, lobby
20

High to
Catastrophic

50

Standoff (ft)
Moderate

limit

Moderate to Low

RISK

High to Moderate

Source: Federal Emergency Management Agency. Primer for Design of Commercial Buildings to Mitigate Terrorist Attacks. FEMA 427 (Washington, DC:
Federal Emergency Management Agency, 2004)

the design enhancements. Using this type of logic, it
is likely to lead to a selection process in which buildings stratify into two groups: those that incorporate
no measures at all or only the most minimal provisions
and those that incorporate high levels of protection. It
also leads to the conclusion that it may not be appropriate to consider any but the most minimal measures
for most buildings.

5.6.14 R
 eferences
In addition to the publications referenced in the text
or figures, the following publications are useful. A
number of the governmental publications may be for
official use only with restricted access.
1 The American Institute of Architects. 2001. Building
Security Through Design: A Primer for Architects,

524

|

ARCHITECTURAL PRECAST CONCRETE

Design Professionals, and Their Clients. Washington
DC: The American Institute of Architects.
2. Mays, G. C., and P. D. Smith. 1995. Blast Effects on
Buildings: Design of Buildings to Optimize Resistance
to Blast Loading. London: Thomas Telford, Ltd.
3. American Society of Civil Engineers. 1997. Design
of Blast Resistant Buildings in Petrochemical
Facilities. Reston, VA: ASCE.
4. U.S. Department of State, Bureau of Diplomatic
Security. March 1998. Architectural Engineering
Design Guidelines (5 Volumes). Washington, DC.
5. U. S. Department of State, Bureau of Diplomatic
Security. August 1995. Structural Engineering
Guidelines for New Embassy Office Buildings.
Washington, DC.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.6.14 References / 5.7.1 Fire Resistance General

6. Krauthammer, T. 2004. Conventional Blasts, Ballistic
Attack, and Related Threats. The Construction
Specifier (May).
7. SBEDS (SDOF Blast Effects Design Spreadsheets).
SBEDS is an Excel© based tool for the design and
analysis of structural components subjected to
dynamic loads, such as air blast from explosives,
using single degree of freedom (SDOF) methodology. SBEDS is based on Army TM 5-1300 (also
designated as NAVFAC P-397 and AFR 88-22) and
UFC 3-340-01 but draws on other sources where
improved methodologies are available. Download
of SBEDS by U.S. government agencies and their
contractors is available from the Protective Design
Center website https://pdc.usace.army.mil/software/sbeds. All other requests for SBEDS should be
emailed to DLL-CENWO-PDC-HD@nwo02.usace.
army.mil.
8. CEDAW (Component Explosive Damage Assessment
Workbook). CEDAW is an Excel-based tool for
generation of pressure-impulse (P-i) and charge
weight-standoff (CW-S) damage-level curves for
structural components. The U.S. Army Corps of
Engineers Protective Design Center developed
CEDAW as a tool for designers to use in satisfying
Department of Defense (DoD) antiterrorism standards. Download of CEDAW by U.S. government
agencies and their contractors is available from the
PDC website https://pdc.usace.army.mil. All other
requests for CEDAW should be emailed to dll-cenwo-PDC-CEDAW@nwo02.usace.army.mil.
9. Protective Structures Automated Design System
(PSADS). Request through the headquarters of the
U.S. Army Corps of Engineers; Attention: CEMPET, 441 G. Street NW, Washington, DC 203141000. PSADS includes DAHS, BlastX, and SPAn32.
Restricted distribution to government agencies and
their contractors. If you are a government contractor, request government project manager to obtain
the software.

5.7 F IRE RESISTANCE
5.7.1 G
 eneral
In the interest of life safety and property protection,
building codes require that resistance to fire be considered in the design of buildings. The degree of fire resistance required depends on the type of occupancy, the

5

size of the building, its location (proximity to property
lines and within established fire zones), and in some
cases, the amount and type of fire detection and suppression equipment available in the structure. Precast
concrete members are inherently non-combustible and
can be designed to meet any degree of fire resistance
that may be required by building codes, insurance
companies, and other authorities.
In Fig. 5.7.1, the dimensional constraints imposed by
the site required building to the property lines. A precast concrete panel system was selected over a unit
masonry wall system to cost-effectively solve the problem of the required 4-hour, fire-rated, exterior propertyline walls that are architecturally consistent with street
elevations. These large walls had to be considered as
“temporary” with the prospect of being concealed by
adjacent buildings some time in the future.
Although life safety is of paramount importance, casualty insurance companies and owners are also concerned with the damage that might be inflicted on the
building and its contents during a fire. This means that
both fire containment and fire resistance must be considered. Insurance rates are often substantially lower
for buildings with higher fire-resistance ratings and
containment designs. In the past, fire-resistance ratings
were assigned on the basis of results of standard fire
tests. In recent years, there has been a trend toward
calculating the fire endurance of building components,
rather than relying entirely on fire tests. To facilitate
this trend, much research work has been conducted on
the behavior of materials and building components in
fires. This section summarizes the available information
on the behavior of architectural precast concrete under
fire conditions. See Section 4.5.7 for a discussion on
fire protection of connections and Section 4.7.9 for fire
resistance of joints.
Fire resistance ratings of building components are
measured and specified in accordance with a common
standard, ASTM E119. Fire endurance is defined as the
period of time elapsed before a prescribed condition
of failure or end point is reached during a standard fire
test. The major “end points” used to evaluate performance in a fire test include:
1. Collapse of loadbearing specimens (structural end
point).
2 	Formation of holes, cracks, or fissures through
which flames or gases hot enough to ignite cotton
waste may pass (flame passage end point).

ARCHITECTURAL PRECAST CONCRETE

|

525

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.1 Fire Resistance General

Fig. 5.7.1
832 Folsom Street, San Francisco, California;
Architect: Patri-Merker Architects formerly Whisler-Patri;
Photo: Patri-Merker Architects.

3. T emperature increase of the unexposed surface of
floors, roofs, or walls reaching an average of 250
°F (122 °C) or a maximum of 325 °F (163 °C) at
any one point (heat transmission end point).

526

|

ARCHITECTURAL PRECAST CONCRETE

4. 	Collapse of walls and partitions during a hosestream test or inability to support twice the superimposed load following the hose-stream test.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.7.1 Fire Resistance General / 5.7.2 Fire Endurance of Walls

5

Table 5.7.1. Fire Safety Provisions in IBC 2006 Edition,* Provisions to Prevent Building-to-Building Fire Spread.

Exterior walls
(Load bearing – ratings)

Tables 601 & 602 – Rating depends on type of construction but not less than required
for exterior walls based on fire separation distance, type of construction, and building
use.

Exterior walls
(Non-load bearing
– ratings)

Table 602 – Ratings vary based on setback distances, type of construction, and
building use, with zero rating allowed for setbacks >30 ft. For setbacks >5 ft, rating
required for inside exposure only.

Exterior walls
(Parapets)

704.11 – Extend 30 in. above roof except where:
1. Exterior walls don’t require ratings.
2. Roofs terminate at 2-hr noncombustible roofs.
3. Roofs terminate at 1-hr combustible roofs sheathing with Class B roofing.

Exterior walls
(Limitations on area
of openings)

704.8 – Different percent limits for unprotected and protected openings based on
setback distances, with unlimited openings allowed over 30 ft and 20 ft, respectively.
In other than H-1, H-2, H-3 occupancies, allows opening percent for unprotected to
equal protected for sprinklers.

Combustibility of
cladding on exterior
walls

1406.2 & 2603.5 – Combustible exterior wall finish, other than Fire Resistant Treated
Wood, limited to 10% of wall area where less than or equal to 5 ft setback distance
is provided. On exterior walls of Types I, II, III, and IV construction, combustible trim is
not permitted more than 3 stories or 40 ft above grade plane. Neither of the above
applies to foam plastic complying with Section 2603.6.

*One- and two-family dwelling and townhouse provisions are not a part of this analysis.

A fire-resistance rating (sometimes called a fire rating, a fire-resistance classification, or an hourly rating)
is a legal term defined in building codes, usually based
on fire endurance. Building codes specify required
fire-resistance ratings for various types of construction, occupancy, and fire separation distance. Table
5.7.1 shows fire safety provisions relative to walls in
the IBC 2006 edition. Performance is defined by the
authorities (regulatory and insurance) as the maximum time for which each component would survive
if it were subjected to a standard test. The standard
tests provide arbitrary fire exposure, arbitrary load,
and arbitrary restraint.

5.7.2 F ire Endurance of Walls
The fire endurances of precast concrete walls, as determined by fire tests, are almost universally governed
by the ASTM E119 criteria for heat transmission (temperature rise of the unexposed surface) rather than
by structural behavior during fire tests. This is probably due to the low level of stresses, even in concrete
bearing walls, and the fact that reinforcement generally does not perform a primary structural function. In
most cases, the amount of concrete cover protection
for structural design exceeds the amount required for
fire protection, and so there is, in effect, reserve structural fire endurance within the concrete wall.

Most of the information on heat transmission was derived from fire tests of assemblies tested in a horizontal
position simulating floors or roofs. The data are slightly
conservative for assemblies tested vertically, that is, as
walls. Nevertheless, it is suggested that no correction
be made unless more specific data derived from fire
tests of walls are used.
For concrete wall panels, the temperature rise of the
unexposed surface depends mainly on the thickness
and aggregate type of the concrete. Other less important factors include unit weight, moisture condition,
air content, and maximum aggregate size. Within the
usual ranges, water-cement ratio, strength, and age
have insignificant effects.
From information that has been developed from fire
tests, it is possible to accurately estimate the thickness
of many types of one-course and multi-course walls that
Table 5.7.2. Fire Endurances for Single-Mixture Concrete Panel.

Thickness for fire endurance, in.
Aggregate

1 hr

2 hr

3 hr

4 hr

All lightweight

2.47

3.56

4.35

5.10

Sand-lightweight

2.63

3.76

4.62

5.37

Carbonate

3.25

4.67

5.75

6.63

Siliceous

3.48

5.00

6.15

7.05

ARCHITECTURAL PRECAST CONCRETE

|

527

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.2 Fire Endurance of Walls

Table 5.7.3. Thickness of Wythes to Provide Various Fire Endurances for Panels with Facing and Backup Materials.

Fire
Endurance,
hr

Backup Material

Siliceous Aggregate
Concrete, in.
(Facing Material)

Sand-Lightweight
Concrete, in.
(Facing Material)

Inside Wythe Material
(Fire-Exposed Side)

11/2

2

3

11/2

2

3

1

Carbonate aggregate concrete*

1.9

1.4

0.45

1.7

1.0

0

1

Siliceous aggregate concrete

2.0

1.48

0.48

1.7

1.0

0

1

Lightweight aggregate concrete

1.5

1.2

0.25

1.13

0.63

0

1

Cellular concrete (30 pcf)

0.7

0.5

0.2

0.5

0.3

0

1

Perlite concrete (30 pcf)

0.8

0.6

0.2

0.7

0.4

0

1

Vermiculite concrete (30 pcf)

0.9

0.6

0.2

0.7

0.4

0

1

Sprayed mineral fiber

0.4

0.25

0.1

0.4

0.2

0

1

Sprayed vermiculite cementitious material

0.4

0.25

0.1

0.4

0.2

0

2

Carbonate aggregate concrete*

3.25

2.8

1.9

3.2

2.6

1.25

2

Siliceous aggregate concrete

3.5

3.0

2.0

3.3

2.7

1.3

2

Lightweight aggregate concrete

2.5

2.1

1.4

2.26

1.76

0.76

2

Cellular concrete (30 pcf)

1.2

1.0

0.6

1.2

0.9

0.4

2

Perlite concrete (30 pcf)

1.4

1.1

0.7

1.3

0.9

0.4

2

Vermiculite concrete (30 pcf)

1.6

1.3

0.8

1.4

1.1

0.4

2

Sprayed mineral fiber

1.1

0.8

0.5

1.0

0.8

0.3

2

Sprayed vermiculite cementitious material

1.0

0.8

0.5

1.0

0.75

0.3

3

Carbonate aggregate concrete*

4.4

3.9

3.0

4.2

3.7

2.4

3

Siliceous aggregate concrete

4.65

4.15

3.15

4.4

3.8

2.5

3

Lightweight aggregate concrete

3.4

3.1

2.4

3.12

2.62

1.62

3

Cellular concrete (30 pcf)

1.6

1.3

0.9

1.6

1.3

0.8

3

Perlite concrete (30 pcf)

1.9

1.6

1.1

1.8

1.4

0.8

3

Vermiculite concrete (30 pcf)

2.2

1.8

1.3

2.0

1.6

1.0

3

Sprayed mineral fiber

NA

1.4

0.9

NA

1.3

0.85

3

Sprayed vermiculite cementitious material

1.6

1.35

0.85

1.6

1.3

0.8

4

Carbonate aggregate concrete*

5.15

4.8

3.85

5.2

4.7

3.5

4

Siliceous aggregate concrete

5.55

5.05

4.05

5.5

4.9

3.7

4

Lightweight aggregate concrete

4.2

3.8

3.0

3.87

3.37

2.37

4

Cellular concrete (30 pcf)

2.1

1.9

1.4

2.0

1.7

1.1

4

Perlite concrete (30 pcf)

2.3

2.0

1.5

2.3

1.9

1.3

4

Vermiculite concrete (30 pcf)

2.7

2.3

1.7

2.6

2.2

1.5

4

Sprayed mineral fiber

NA

NA

1.4

NA

NA

1.4

4

Sprayed vermiculite cementitious material

NA

1.8

1.3

1.75

1.75

1.25

*Tabulated values for thickness of inside wythe are conservative for carbonate aggregate concrete.
Note: 1. NA = not applicable; that is, a thicker facing material is needed.
2. To obtain thickness of concrete for a specific fire endurance, read across and then up. For example, a 2 hr fire endurance for a 2 in. siliceous facing and
carbonate backup requires 4.8 in. of concrete.

528

|

ARCHITECTURAL PRECAST CONCRETE

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.2 Fire Endurance of Walls

Fig. 5.7.2 Fire endurance (heat transmission) as a function
of panel thickness.

Table 5.7.4 Use of 5/8 in. Type X Gypsum Wallboard.

Thickness of Concrete Panel
for Fire Endurance, in.

5

With 7/8 in.
air space

Lightweight — 100 PCF

4

Fire Endurance, Hr

Sand-Lightweight — 115 PCF

3

Carbonate Aggregate

0

3

4

5

Aggregate

2 hr

3 hr

2 hr

3 hr

Sand-lightweight

2.5

3.6

2.0

2.5

Carbonate

2.8

4.0

2.0

2.7

Siliceous

2.9

4.2

2.0

2.8

4. Siliceous aggregates include quartzite, granite,
basalt, and most rocks other than limestone and
dolomite.

Siliceous Aggregate

2

With 6 in.
air space

3. Carbonate aggregates include limestone and dolomite (minerals consisting mainly of calcium and/or
magnesium carbonate).

2

1

5

6

7

Wall Panel Thickness, In.

will provide fire endurances of 1, 2, 3, or 4 hours, based
on the temperature rise of the unexposed surface (refer
to ACI 216.1/TMS 0216.1). Based on fire test data, the
thicknesses shown in Fig. 5.7.2 and Tables 5.7.2 and
5.7.3 can be expected to provide the fire endurances
indicated for single-course and two-course walls. Figure
5.7.2 shows the fire endurance (heat transmission) of
concrete as influenced by aggregate type and thickness.
Interpolation of varying concrete unit weights is acceptable in this figure. Table 5.7.2 provides the thickness (in
inches) of solid concrete wall panels for various fire endurances, while Table 5.7.3 provides the same for twocourse panels.
As used in this section, concrete aggregates are designated as lightweight, sand-lightweight, carbonate,
or siliceous.
1. L ightweight aggregates include expanded clay,
shale, slate, and sintered fly ash. These materials
produce concretes having unit weights of about
95 to 105 pcf (1520 to 1680 kg/m3) without sand
replacement.
2. Lightweight concretes in which sand is used as part
of or all of the fine aggregate, and unit weight of
105 to 120 pcf (1680 to 1920 kg/m3), are designated as sand-lightweight.

Ribbed panel heat transmission is influenced by
both the thinnest portion of the panel and by the panel’s “equivalent thickness.” Here, equivalent thickness
is defined as the net cross-sectional area of the panel
divided by the width of the cross-section. In calculating
the net cross-sectional area of the panel, portions of
ribs that project beyond twice the minimum thickness
should be neglected (Fig. 5.7.3).
The fire endurance (as defined by the heat transmission end point) can be governed by either the thinnest
section, the average thickness, or a combination of the
two. The following rule-of-thumb expressions describe
the conditions under which each set of criteria governs.
Let t

= minimum thickness, in.

te = equivalent thickness of panel, in.
s

= rib spacing, in.
s
If t ≤ , fire endurance, R, is governed by t and is
4
equal to Rt.
s
If t ≥ 2 , fire endurance, R, is governed by te and
is equal to Rte.
s
s
If > t > :
2
4
 4t 
R = R t +  - 1 R te - R t
 s


(

)

(Eq. 5.7.1)

where R is the fire endurance of a concrete panel and
subscripts t and te relate the corresponding R values to
concrete slab thicknesses t and te, respectively.

ARCHITECTURAL PRECAST CONCRETE

|

529

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.2 Fire Endurance of Walls

Fig. 5.7.3 Cross section of ribbed wall panels.

2t

t

Neglect Shaded Area
in Calculation of
Equivalent Thickness
(a)

te

t

Window walls that are required to be fire resistive
have limits imposed on the area of openings by the
building code. These limits are based on the construction classification, occupancy, spatial separation (distance between a building and its neighbor or property
line), and fire zone. For example, Table 704.8 of the

s
(b)

These expressions apply to ribbed and corrugated panels, but they give excessively low results for panels with
widely spaced grooves or rustications. Consequently,
engineering judgment must be used when applying
the above expressions.
Sandwich panels have insulating materials between the two wythes of concrete (see Section 5.3.8).
Building codes require that, where non-combustible
construction is specified, combustible elements in walls
be limited to thermal and sound insulation having a
flame-spread index of not more than 100, except it
should not exceed 75 for foam plastic insulation when
the insulation is sandwiched between two layers of
non-combustible material such as concrete. When insulation is not installed in this manner, it is required to
have a flame-spread index of not more than 25. Data
on flame-spread classification are available from insulation manufacturers.
It should be noted that the cellular plastics melt and
are consumed at about 400 to 600 °F (205 to 316
°C). Thus, thickness greater than 1.0 in. (25 mm) or
changes in composition probably have only a minor
affect on the fire endurance of sandwich panels. The
danger of toxic fumes caused by the burning of cellular
plastics is practically eliminated when the plastics are
completely encased within concrete sandwich panels.
It is possible to calculate the thicknesses of various materials in a sandwich panel required to achieve a given
fire rating using Equation 5.7.2.
R0.59 = R10.5 9 + R20.5 9 … Rn0.59

(Eq. 5.7.2)

where R = fire endurance of the composite assembly
in minutes and R1, R2, and Rn = fire endurance of each of the individual courses in
minutes.

530

|

ARCHITECTURAL PRECAST CONCRETE

Table 5.7.5 lists fire endurances for insulated precast
concrete sandwich panels with either cellular plastic,
glass fiber board, or insulating concrete used as the
insulating material. The values were obtained using Eq.
5.7.2. A design graph for solving the equation is provided (Fig. 5.7.4).

Table 5.7.5 Fire Endurance of Sandwich Panels.

Inside
Wythe,
in.

Fire
Insulation,
Outside Endurance,
in.
Wythe, in.
hr:min

11/2 Sil

1 CP

11/2 Sil

1:23

1 /2 Carb.

1 CP

1 /2 Carb.

1:23

1 /2 SLW

1 CP

1 /2 SLW

1:45

2 Sil

1 CP

2 Sil

1:50

2 Carb.

1 CP

2 Carb

2:00

2 SLW

1 CP

2 SLW

2:32

1

1

3 Sil

1

1

1 CP

3 Sil

3:07

11/2 Sil

3

/4 GFB

11/2 Sil

1:39

2 Sil

3

/4 GFB

2 Sil

2:07

2 SLW

3

/4 GFB

2 SLW

2:52

2 Sil

3

/4 GFB

3 SLW

3:10

1

1 /2 Sil

1 /2 GFB

1

1 /2 Sil

2:35

2 Sil

1 /2 GFB

2 Sil

3:08

2 SLW

1 /2 GFB

2 SLW

4:00

1
1
1

1 /2 Sil

1 IC

1 /2 Sil

2:12

11/2 SLW

1 IC

11/2 SLW

2:39

2 Carb

1 IC

2 Carb

2:56

2 SLW

1 IC

2 SLW

3:33

1 /2 Sil

1 /2 IC

1 /2 Sil

2:54

1 /2 SLW

1 /2 IC

1 /2 SLW

3:24

2 Sil

1 /2 IC

3 Sil

4:16

2 Sil

2 IC

2 Sil

4:25

1 /2 SLW

2 IC

1 /2 SLW

4:19

1

1

1

1

1
1

1

1

1

1

1

Note: Carb = carbonate aggregate concrete; Sil = siliceous aggregate
concrete; SLW = sand-lightweight concrete (115 pcf maximum); CP = cellular
plastic (polystyrene or polyurethane); GFB = glass fiber board; IC = lightweight insulating concrete (35 pcf maximum).

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.2 Fire Endurance of Walls

5

Fig. 5.7.4 Design aid for use in solving Eq. 5.7.2.

10

Co
nc
Lig
re
te
ht
we
igh
tC
on
cre

35
—

R0.59

60

11.20

120

16.85

180

21.41

te

eig

nd
Sa

tw
Lig
h

n

rbo

Ca

us

eo

ic
Sil

ate

g

Ag

240

25.37

te

a
reg

Material

R0.59

cellular plastic (1 in. or thicker)

2.5

/4 in. glass fiber board

4.0

1

1 /2 in. glass fiber board

8.5

continuous air space

3.33

te

cre

a
reg

ht

Co
ncr
ete
ing
Ins
ula
t

R0.59 for use in Equation 5.7.2

15

R1 Minutes
te

PC
F

20

n
Co

te

e
ncr

Co

g
Ag

3

5

two continuous air spaces

6.67

2 in. foam glass

10.6

/2 in. gypsum wallboard

7.44

/8 in. gypsum wallboard

8.49

1

0

5

1

2

3

4

5

Equivalent Thickness of One Course, In.

IBC permits no unprotected openings in exterior walls
when the spatial separation is less than 5 ft (1.5 m) or
less than 3 ft (0.9 m) for protected openings. Exterior
protected openings (covered in IBC Section 715) required to have a fire-resistance rating of greater than
1 hour are to be protected with an assembly having a
fire-protection rating of not less than 1.5 hours. If required fire-resistance rating is 1 hour, then the assembly can have a fire-protection rating of not less than
45 minutes.
For buildings that are three stories or more in height,
openings in exterior walls in adjacent stories are to be
separated vertically to protect against fire spread on
the exterior of the buildings where the openings are
within 5 ft (1.5 m) of each other horizontally and the
opening in the lower story is not a protected opening.
Such openings are to be separated vertically by at least
3 ft (1 m) by spandrel girders, exterior walls, or other
similar assemblies that have a fire-resistance rating of
at least 1 hour or by flame barriers having a fire-resistance rating of at least 1 hour that extends horizontally
at least 30 in. (762 mm) beyond the exterior wall.

Requirements for various occupancies differ somewhat but generally follow the same pattern and certain exceptions often apply. The IBC code relates
spatial separation and maximum area of unprotected
openings to the area of the exposed building face.
Percentages of unprotected opening areas are then
tabulated in the code for various combinations of area
of building face and spatial separation. The percentage
of openings permitted increases as the spatial separation increases. IBC also permits a higher limit on the
unexposed surface temperature if the area of unprotected openings is less than the maximum allowed
and, thus, somewhat thinner panels can be used.
An equivalent opening factor, Feo, is then applied in a
formula to determine the corrected area of openings:
Ae = A + AfFeo

(Eq. 5.7.3)

where
Ae = equivalent area of protected openings.
A = actual area of protected openings
Af = area of exterior surface of the exposed building

ARCHITECTURAL PRECAST CONCRETE

|

531

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.2 Fire Endurance of Walls / 5.7.3 Detailing of Fire Barriers

face exclusive of openings, on which the temperature limitation of the standard fire test is
exceeded.

an opening which would occur with complete through
penetration.
1. Unless the wall is very close to the property line,
openings are permitted. If protection is needed,
then 45 minutes is all that is required. It will take
only about 3 in. (75 mm) of concrete to get a 45minute rating, so this should present no problem.

Figure 5.7.5 shows the relation between Feo and panel
thickness for three types of concrete.
To illustrate the use of Fig. 5.7.5, suppose that for
a particular building face, a 2-hour fire-resistance rating is required and the area of unprotected openings
permitted is 57%. Suppose also that the actual area
of unprotected openings is 49% and that the window
wall panels are made of carbonate aggregate concrete.
Determine the minimum thickness of the panel.

2. If the wall is far enough off property line, 5 ft (1.5
m) or more, so that opening protection is not required, nothing needs to be done. Table 704.8 of
the IBC allows for unprotected openings not to
exceed 10% of the total wall area, a number that
would never be realized.

In this case: Ac = 57%,

3. If the wall is close to the property line, within 5 ft
(1.5 m), then openings may not be allowed. This
could require restoring the pocket to its original
fire rating by applying a fire-resistive spray, inserting various fire-retardant materials, or moving the
pocket to the opposite wall (using a corbel or ledge
instead of a pocket).

A = 49%, Af = 100-49 = 51%,
hence:
Feo =

A c - A 57 - 49
=
= 0.16
Af
51

From Fig. 5.7.5, for Feo = 0.16 at two hours, the minimum panel thickness is 2.3 in. (33 mm). Thus, if the
panel is 2.3 in. (33 mm) thick or thicker, the code requirements will be satisfied.

5.7.3 Detailing of Fire Barriers
One of the purposes of code provisions for fire-resistive construction is to limit the involvement of a fire to
the room or compartment where the fire originates.
Thus, the floors, walls, and roof surrounding the compartment must serve as fire barriers.

When a horizontal member pockets into a wall, reducing thickness (t) by half, there are several ways to
approach the fire rating of the wall. One approach is
to consider the reduction in wall thickness the same as

Fig. 5.7.5 Equivalent Opening Factor, Feo.

Feo Equivalent Opening Factor

Carbonate Aggregate
Concrete (Type N)

Siliceous Aggregate
Concrete (Type S)

0.4

0.4

0.4

0. 3

0. 3

0. 3

0. 2

2 hr

0. 2

2 hr
3/4

0. 1

0. 2

1

1

1

3/4

0. 1

0. 1

1

2

3

4

5

0

1

2

3
Panel Thickness, In.

|

2 hr

3/4

0

532

Sand-Lightweight
Concrete (Type LS)

ARCHITECTURAL PRECAST CONCRETE

4

5

0

1

2

3

4

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.7.3 Detailing of Fire Barriers / 5.7.4 Columns and Column Covers

IBC requires that firewalls start at the foundation and
extend continuously through all stories to and above
the roof, except where the roof is of fire-resistive construction, in which case the wall must be tightly fitted
against the underside of the roof. If the roof and walls
are of combustible construction, firewalls must extend
not only through the roof, but must extend through
the sides of the building beyond the eaves or other
combustible projections.
When precast concrete wall panels are designed and
installed such that no space exists between the wall
panel and floor, a fire below the floor cannot pass
through the joint between the floor and wall. However,
some wall panels are designed such that a space does
exist, a space referred to as a “safe-off” area.
Figure 5.7.6 shows a method of fire stopping such
safe-off areas. Safing is supported on a steel angle,
with or without Z-shaped impaling pins, depending
on gauge of steel angle. Safing insulation is available as mineral fiber mats of varying dimensions and
densities.
The mineral fiber should be sealed with a firestop
caulk. Care must be taken during installation to ensure
that the entire safe-off area is sealed. The safing insulation provides an adequate firestop and accommodates
differential movement between the wall panel and the
floor. See Section 5.5.8 for a discussion on the acoustical isolation effects of this treatment.

5.7.4 C olumns and Column Covers
Reinforced precast concrete columns have for many
years served as the standard for fire-resistive construction. Indeed, the performance of concrete columns in
actual fires has been excellent.
The inherent fire resistance of concrete columns results from three factors:
1. Minimum size of a structural column is generalTable 5.7.6. Minimum Sizes of Concrete Columns.

Minimum Column Size for
Fire Resistance Rating, in.
Aggregate type

1 hr 11/2 hr

2 hr

3 hr

4 hr

Siliceous

8

9

10

12

14

Carbonate

8

9

10

11

12

Sand-Lightweight

8

8 /2

9

10 /2

12

1

1

5

Fig. 5.7.6 Methods of installing safing insulation.

Sealant
Safing Insulation
Continuous Plates

Metal Plate Attached to Wall
1 / " to 8" Maximum
1 2

Wall Panel

ly such that the inner core of the column retains
much of its strength even after long periods of fire
exposure.
2. Concrete cover to the primary reinforcing bars is
generally 17/8 in. or more, thus providing considerable fire protection for the reinforcement.
3. Ties or spirals contain the concrete within the
core.
Table 5.7.6 shows typical building code requirements
for reinforced concrete columns, and the values shown
apply to both precast and cast-in-place concrete columns. In addition, they apply to cast-in-place concrete
columns clad with precast concrete column covers,
whether the covers serve merely as cladding or as
forms for the cast-in-place column.
Precast concrete column covers are often used to
clad steel columns for architectural reasons. Such covers also provide fire protection for the columns. Figure
5.7.7 shows the relationship between the thickness of
a concrete column cover and the fire endurance for
various steel column sections. The fire endurances
shown are based on an empirical relationship. It was
also found that the air space between the steel core
and the column cover has only a minor affect on the
fire endurance. An air space will probably increase the
fire endurance but by an insignificant amount.
Most precast concrete column covers are 3 in. (75
mm) or more in thickness, but some are as thin as 21/2
in. (63 mm). From Fig. 5.7.7, it can be seen that such
column covers provide fire endurances of at least 2.5

ARCHITECTURAL PRECAST CONCRETE

|

533

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.4 Columns and Column Covers

Fig. 5.7.7 Fire endurance of steel columns afforded protection by concrete column covers.
5

5

4

4
W14 x 320

W14 x 320
W10 x 112

W10 x 112
Fire Endurance, Hr

3

3
W8 x 31

W8 x 31

W10 x 49

W10 x 49

2

2
t

1

1
Normal Weight Concrete
(Carbonate or Siliceous Aggregates)

0

0

1

2
(a)

3

4

(b)

(c)

For example, the fire endurance afforded by a 3 in.thick (75mm) column cover of normalweight concrete
for an 8 × 8 × 1/2 in. (200 × 200 × 13 mm) steel tube
column will be about 3 hours 20 minutes (the weight
of the section is 47.35 lb/ft [691 N/m]).
Figure 5.7.8 displays some of the various shapes of

|

ARCHITECTURAL PRECAST CONCRETE

0

1

2
(b)

3

4

precast concrete column covers, including (a) two
L-shaped units, (b) two mitered units, and (c) two
U-shaped units. There are, of course, many other combinations that may be used to accommodate isolated
columns, corner columns, and columns in walls.

hours and usually more than 3 hours. For steel column
sections other than those shown, including shapes
other than wide flange beams, interpolation between
the curves on the basis of weight per foot will generally give reasonable results.

534

0

T1 Thickness of Column Cover, In.

Fig. 5.7.8 Types of column covers.

(a)

Sand-Lightweight — Concrete

To be fully effective, the column covers must remain
in place without severe distortion. Many types of connections are used to hold the column covers in place.
Some connections consist of bolted or welded clip angles attached to the tops and bottoms of the covers.
Others consist of steel plates embedded in the covers that are welded to angles, plates, or other shapes
which are, in turn, welded or bolted to the steel column. In any case, these connections are used primarily to position the column covers and as such, are not
highly stressed. As a result, temperature limits do not
need to be applied to the steel in most column cover
connections.
If either partially or fully restrained, concrete panels
tend to deflect or bow when exposed to fire. For example, for a steel column that is clad with four flat
panels attached top and bottom, the column covers

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS

5.7.4 Columns and Column Covers / 5.7.6 Protection of Connections

will tend to bulge at mid-height, opening gaps along
the sides. The gap sizes decrease as the panel thicknesses increase.
With L-, C-, or U-shaped panels, the gap size is further reduced. The gap size can be further minimized
by connections installed at mid-height. In some cases,
shiplap joints can be used to minimize the effects of
joint openings.
Joints should be sealed to prevent the passage of
flame to the steel column. A non-combustible material, such as sand-cement mortar, or a ceramic-fiber
blanket can be used to seal the joint and then caulking
is applied.
Precast concrete column covers should be installed in
such a manner that if they are exposed to fire, they
will not be restrained vertically. As the covers are heated, they tend to expand, and the connections should
accommodate this expansion without subjecting the
cover to additional loads. For this reason, the precast
concrete column covers should not be restrained vertically. Fire-resistive compressible materials, such as mineral fiber safing, can be used to seal the tops or bases
of the column covers, permitting the column covers
to expand without restraint. Similarly, the connections
between the covers and columns should be flexible
(or soft) enough to accommodate thermal expansion
without inducing much stress into the covers.

5.7.5 P
 rotection of Reinforcing Steel
For the purpose of establishing fire ratings, the codes
currently do not address cover for the reinforcing steel
in walls, an oversight that will, perhaps, be remedied
in the near future. The codes do require that the serviceability requirements be met by providing adequate
cover for protection against weather and other effects.
Because of these precautions and the proven performance of concrete walls in real fires, it appears that
these requirements furnish the necessary cover demanded for fire ratings. It is recommended that for fire
ratings of 1 hour through 4 hours, concrete cover be
furnished as specified in ACI 318, Section 7.7.3. It is
also recommended that regardless of the type of aggregate used in the concrete, the minimum thickness
of concrete cover to the main longitudinal reinforcement should not be less than 1 in. (25 mm) times the
number of hours of required fire resistance, or 2 in. (50
mm), whichever is more.

5

5.7.6 Protection of Connections
Fireproofing of connections may be necessary, depending on codes and/or insurance requirements. In
many cases, fireproofing with concrete cover will also
provide corrosion protection. Many types of connections in precast concrete construction are not vulnerable to the effects of fire and, consequently, require
no special treatment. For example, direct bearing areas between precast concrete panels and footings or
beams that support them do not generally require any
special fire protection, nor do concrete haunches.
If the panels rest on elastomeric pads or other combustible materials, protection of the pads is not generally required because pad deterioration will not cause
collapse. Nevertheless, after a fire, the pads would
probably have to be replaced, so protecting the pads
might prevent the need for replacement. If the connections are to be fireproofed or concealed, this fact
should be indicated in the contract documents.
Connections that can be weakened by fire and thereby
jeopardize the structure’s load-carrying capacity should
be protected to the same degree as that required for
the structural frame. If, for example, when an exposed
steel bracket supports a precast concrete element that
is required to have a designated fire rating, the steel
bracket must be protected to the same fire rating.
Many connections simply provide stability and are under little or no stress in service. While fire could substantially reduce the strength of such a connections, no fire
protection is necessary. Connections that have steel elements encased in concrete, drypacking, or grout after
erection usually need no additional protection.
There is evidence that exposed steel hardware used in
connections is less susceptible to fire-related strength
reduction than other steel members. This is because
the concrete provides a “heat sink,” which draws off
the heat and reduces the temperature of the steel.
Fireproofing of connections is usually accomplished
with sprayed cementitious or mineral fiber fireproofing, intumescent mastic compounds, or enclosure with
gypsum wallboard.
Figure 5.7.9(a) shows the thicknesses of various,
commonly used, fire-protection materials required for
fire endurances up to four hours when applied to connections consisting of structural steel shapes. The values shown are based on a critical steel temperature

ARCHITECTURAL PRECAST CONCRETE

|

535

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.7.6 Protection of Connections / 5.8.2 Roofing Flashing

Fig. 5.7.9(a) & (b) Thickness of protection materials applied to connections consisting of structural steel shapes.
(IM = intumescent mastic, SMF = sprayed mineral fiber, VCM = vermiculite cementitious material).
Concrete or Dry-Pack Mortar
b = minimum width of concrete protection

4

4

3

3

F

t=

CM

8"

"V

5 /16

2

b≥

CM
or V
SMF 11 /16"
t=

1

0

SM

=

t=

2

5

"
/16

2

3

12"

1

IM
"
t = 5 /16
1

b

Thickness of Protection Material, in.

t = minimum thickness of steel subjected
to fire from both sides

0

4

1

2

3

4

Fire Endurance, hr
(a)

of 1000 °F (538 °C) (that is, a stress-strength ratio of
about 65%). The values in Fig. 5.7.9(b) are applicable
to concrete or drypack mortar encasement of structural steel shapes used as brackets.

the installation method used, and the quality of the
work. Refer to the NRCA Construction Details in the
NRCA Roofing and Waterproofing Manual for industry
accepted recommended practices.

When a rational analysis or design for fireproofing is
not performed and concrete is used to fireproof the
connections in the field, a conservative estimate would
suggest that such concrete should have a thickness
in inches corresponding to the specified hours of fire
rating. Unless the nature of the detail itself supports
such concrete, it should be reinforced with a light wire
fabric.

Three major considerations govern the detailing of
roofing adjoining precast concrete units:

5.8 R
 OOFING
5.8.1 G
 eneral
The most vulnerable parts of any roof system, relative to leaks, are the interfaces between the horizontal roof surface and vertical surfaces and penetrations.
Therefore, designers should carefully consider the design of flashing details at these locations and at system
terminations. Flashings are as important to the performance of a low-slope roof as the membrane selected,

536

|

(b)

ARCHITECTURAL PRECAST CONCRETE

1. Relative movement between the roof deck and the
precast concrete panels should be assessed by the
designer, and allowances made to accommodate
this movement with flashing, counterflashing, and
expansion joints.
2. Hardware required in the precast concrete should
be detailed and located to tolerances stated on the
working drawings and selected to suit the particular conditions.
3. Details should reflect involvement of the minimum
number of trades, ensuring that the work of each
trade can be completed independently of the
others.

5.8.2 Flashing
The basic elements of flashing design apply to all types

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.2 Roofing Flashing

5

Fig. 5.8.1 Counter flashing for concrete wall or parapet.
Install Appropriate Sealant (e.g., Polyurethane) and Tool
to Facilitate Water Run-off
Concrete Fasteners
Sheet Metal Counterflashing
Optional: Install Compressible Elastomeric Sealant or Tape
to Span Irregularities
Termination Bar Fastened with Appropriate Fasteners
approx. 12" [300 mm] O.C.
Install Compatible Sealant Behind Membrane
Flashing Membrane (8" [300 mm] min. height)
Reinforced Membrane Attachment Strip with Seam Plates
and Fasteners

Thermoset Roof Membrane
Coverboard and Thermal Insulation
Roof Deck

of roofs. Base flashing is a continuation of the roofing membrane that is typically applied separately from
the field application. The flashing components, made
of roof membrane and sheet metal, include cap flashing or counterflashing, which is applied to shield the
exposed portions of the top edge of base flashing or to
extend into the wall to divert any interior water to the
exterior of the wall. Metal copings are the preferred
method to seal and waterproof the top of a parapet or
building wall. Edge flashings (gravel stops) are used
to terminate roofing membrane and hold the gravel in
place on a ballasted roof or to finish off the edge of
other types of roof. Expansion joints are structural
separations that accommodate movement between
two building elements, or at specific locations, such
as where the roof deck changes direction. Roof scuppers provide an exit for water through a parapet wall
or an elevated edge. Each of these flashing components requires equal design consideration.
The details in this section depict jobsite fabricated
construction. Many roofing material manufacturers
now offer prefabricated flashing pieces or permit the
use of materials for flashing purposes other than those

shown here. Specifics on these proprietary designs vary
greatly. Therefore individual roofing material manufacturers should be consulted when proprietary designs
are used. Terminations of built-up roofing, which do
not allow for movements, such as spandrel beam and
deck deflections, at horizontal/vertical transitions can
lead to water entry paths at the roof/wall panel interface. A common detail for application of a single-ply
roof is shown in Fig. 5.8.1. The figure indicates only
the edge of the roof and refers only to materials directly affecting the precast concrete details.
Single-ply membrane systems have evolved along several lines. Thermoplastic sheets such as polyvinyl chloride (PVC) and thermoplastic olefin (polyolefin) (TPO)
are typically internally reinforced with a scrim or fabric
and are generally seamed by hot-air welding. Ethylene
propylene diene monomer (EPDM) membranes are
thermoset sheets, which are often reinforced and the
seams glued or taped. When single-ply membranes
are returned vertically on the back of precast concrete
panels to act as flashing, the back surface of the panel
must be given a smooth trowel finish. Surface irregularities may puncture or tear the roofing membrane.

ARCHITECTURAL PRECAST CONCRETE

|

537

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.2 Roofing Flashing

Fig. 5.8.2(a – c) Typical roof flashing details.
(a)

Appropriate Sealant Tooled to Facilitate Run-off
Appropriate Fasteners 12" [300 mm] O.C.
Surface Mounted Sheet Metal Receiver
Appropriate Fasteners 24" [600 mm] O.C.
3" [75 mm] Lap at Joints
Removable Sheet Metal Counterflashing

(c)

Optional: 2" [50 mm] Wide Clip approx. 30"
[750 mm] O.C. – 2 Fasteners per Clip

Extruded Butyl
Rubber Surface

Membrane Base Flashing
Optional: Sealant or Elastomeric Tape
Two-Piece Surface Mounted Reglet and Counterflashing

(b)

Appropriate Sealant Tooled to
Facilitate Run-off
Appropriate Fasteners 12" [300 mm] O.C.

Reglet is
Fastened
with any
Suitable
Fastener

One-piece Sheet Metal Counterflashing
3" [75 mm] Lap at Joints
Optional: 2" [50 mm] Wide Clip approx.
30" [750 mm] O.C. – 2 Fasteners per Clip
Membrane Base Flashing
Optional: Sealant or Elastomeric Tape

One-Piece Surface Mounted Counterflashing

538

|

Cant strip not
to scale

Alternate Method of Attaching Counterflashing

The bending radius of built-up or modified roofing
materials is generally limited to 45 deg. To allow for
this bending radius, all vertical surfaces must have
cant strips installed between the roof and the vertical surface. The base flashing should extend vertically,
from the horizontal plane of the roof, at least 8 in.
(200 mm). Walls requiring protection higher than 8 in.
(200 mm) should receive different moisture-proof detailing. A wood nailer strip fastened to the roof deck
(not the wall) or suitable detail allowing mechanical
fastening of the membrane base flashing must be provided. Since metals have a high coefficient of expansion, metal flashings and panel connections must be
isolated from the roof membrane wherever possible to
prevent thermal movements of the metal from splitting
or tearing the membrane.

stalled on the wall above the base flashing (Fig. 5.8.2).
An alternate method of attaching the counterflashing
is shown in Fig. 5.8.2(c). The design of this detail should
be two-piece (reglet and counterflashing), allowing installation of the counterflashing after the base flashing
is completed. Single-piece installations are difficult to
flash properly. Also, it is difficult to perform re-roofing
and roofing maintenance without deforming the metal
when single-piece installations are used. Sheet metal
should never be used as a base flashing.

For all walls and projections that receive membrane
base flashing, metal counterflashings should be in-

Differential movement of material components at
roof-to-wall junctions can cause wrinkling and buck-

ARCHITECTURAL PRECAST CONCRETE

The roof to wall junction detail is one of the most frequent areas of failure. The major reason is the failure to
accommodate differential movements caused by temperature changes, moisture changes, building frame
movement, and wind. Especially important is the live
load imposed by snow, or retained rainwater.

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.2 Roofing Flashing / 5.8.3 Parapet Details

5

Fig. 5.8.3(a) & (b) Low parapet details.

1" [25 mm] min.Below Blocking

Sheet Metal Fascia Cap
1" [25 mm] min.Below Blocking

Gasket Fasteners
Fasten Base Flashing
approx. 8" [200 mm] O.C.
Ap
p
[10 rox. 4
0m "
m]

Wood Cant Fastened
to Blocking

Multiple-ply Membrane
Base Flashing at Field
of Roof

Multiple-ply Membrane
Stripping Feathered at Field
of Roof
Fasteners 3"
to 6" [75 to
150 mm] O.C.,
staggered

2" [50 mm] min.
2" [50 mm]
min.

Sheet Metal “Gravelstop” set in compatible
roof cement – Prime
Flange before stripping
Felt Envelope (for coal tar and
asphalt types I and II)
Multiple-ply Builtup Roof Membrane

Extend One Base
Flashing Ply Below Blocking
Multiple-ply Built-up
Roof Membrane

Coverboard Insulation
and Thermal Insulation

Coverboard Insulation
Thermal Insulation
Roof Deck

(a)

Wood Nailers
Continuous Cleat

ling, delamination, loss of adhesion, and open seams
in flashing membranes. Flashing membranes subjected
to differential wall-to-roof movement can separate
from their substrate, tear, and become a source for
water entry into the roof system and/or building interior. NRCA recommends that designers create flashing
details that can accommodate separate roof-to-wall
movement when excessive movement between a roof
and wall is anticipated. This condition may occur where
a roof deck is not supported by a wall. Typically, a perimeter expansion joint at the wall or cornice-to-roof
system will accommodate this differential movement.
Some preformed proprietary roofing material shapes
allow direct fastening to precast concrete, even when
moderate movements are expected.
The roofing edge detail must be able to accommodate any relative vertical and/or horizontal movement
between the wall and roof to prevent rupture of membrane roofing. The lateral connection (tieback) from
the wall to the roof member should be a slotted insert
with a threaded flat bar welded or bolted to the roof
member to allow for differential movement.

Flange approx. 4"
[100 mm] min.

(b)

Roof Deck
Wood Nailers
Extended One Ply
or a Separate Membrane Sheet to Below Blocking
Continuous Cleat

The major requirement for a reglet is strict adherence to close tolerances in placing the reglet. The
precaster may overlook this requirement unless clearly
stated in the contract documents (see tolerances given
in Section 4.6.2). Cast-in reglets are often difficult to
align properly due to erection tolerances. When not
properly aligned, they can hinder the proper installation of counterflashing. Therefore, the use of cast-inreglets is generally not recommended because of the
difficulty in maintaining tolerance in panel fabrication
and erection.
The method of fastening the roofing membrane to
the precast concrete units varies with local practices.
Some roofers favor a continuous wood nailer, which
alleviates the need to fasten into concrete. The extra
cost to the precaster to incorporate a wood nailer into
the precast concrete should be reflected in equal or
greater savings in the roofing contract.

5.8.3 Parapet Details
For low parapet walls, the base flashing should be
fastened to a vertical wood upright whose horizontal

ARCHITECTURAL PRECAST CONCRETE

|

539

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.3 Parapet Details

base is attached to the deck only. After the base flashing has been attached to the wood upright, the metal
wall cap flashing may be installed. Then the counterflashing may be attached to the wall cap, extending
down over the top of the base flashing. This method
allows lateral movements of the wall without damage
to the base flashing.
Where gravel stops are used, they should be raised
above the roof surface using tapered cants and wood
blocking (Fig. 5.8.3 [a]). When this is not possible, the
metal flanges for low-profile gravel stops should be set
in mastic on top of the completed roof membrane and
fastened at close intervals to the wood nailer. A flashing
strip is then applied over the gravel stop edge. Interior
drainage of a roof is recommended, therefore edges
should be raised whenever possible. In the single-ply
detail, prefabricated metal is used in place of the angular wood cant (Fig. 5.8.3[b]). The metal is fastened to a
flat 2 × 6 in. (50 × 150 mm) nailer, and the membrane
is fastened to the nailer and then flashed.
Parapet panels and connections should be designed
to withstand window-washing equipment loads in any
direction and at any point along the top edge of the

Fig. 5.8.4 Diagram of parapet profile.

Rain
Sufficient Height to Stop Water
Blowing Over Onto Face
Minimum
Run-Off
Surface

1:4 Slope
8 to 12 in. min.

parapet. Parapet and roof edges should be designed
to avoid run-off from flat roofs onto the building façade. A parapet of sufficient height (8 to 12 in. [200 to
300 mm]) will normally prevent water on the roof from
blowing over the parapet onto the face of the building. The top of the parapet should slope backward toward the roof for its full width and be narrow so that
dirt accumulating on them does not cause streaking on

Fig. 5.8.5 Alternate coping securements.
Sheet Metal Parapet Cap
Sloped to Drain Roof Side

Horizontal
Legs may be
used to Ease
Installation

Extended Base Flashing
Optional if Double Lock standing
Seam Joints are used with the
Metal Parapet Cap

Continuous Cleat

Gasketed Fasteners

Continuous Cleats

OPTION “A”

Continuous Shim

1" [25 mm] min.
Below Blocking

Fasteners approx. 8"
[200 mm] O.C.
Wood Nailer

Intermittent Clip (Min.
2 Fasteners per clip)

Prime Surface as Required
Ap
p
[10 rox. 4
0m "
m]

Optional: Extended
Field Plies above
Head of Cant

Continuous Cleat
OPTION “C”

Multiple-ply Membrane
Base Flashing at Field
of Roof
Multiple-ply Built-up Roof Membrane
Coverboard Insulation
Thermal Insulation

Cant
Roof Deck

540

|

Continuous Cleat

OPTION “B”

(a)

ARCHITECTURAL PRECAST CONCRETE

Gasketed Fasteners
(b)

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.3 Parapet Details

the building face when washed off (Fig. 5.8.4). A continuous tapered shim installed over the wood blocking
is a common method of sloping the top of the wall to
the roof side. Alternatively, a sloped top edge on the
precast concrete panel could be used.
Parapet walls must be adequately protected from
moisture intrusion to prevent deterioration of the wall
and damage to the roof system, building components,
and interior. Copings are used to cover the top of a
parapet wall and provide a weatherproof cap, which
seals and protects the wall from moisture. Metal is
one of the more common materials used for copings.
Common metal types are galvanized or Galvalume
steel®, standard or extruded aluminum, and copper.
Lead coated copper and copper can cause staining of
the face of the wall.
The parapet cap (flashing) should project at least 1
in. (25 mm) beyond the vertical wall surfaces and have
a proper drip edge to throw run-off water clear of the
wall. Projections of less than 1 in. may permit water to
either flow back or be blown back against wall surfaces. It is important that the flashing is installed straight
and level and without gaps to avoid streaking from
water run-off. The choice of flashing material and/or
its treatment against corrosion should be based on
preventing potential staining of the precast concrete
surface (see Section 3.6.3).
When weather-tight joints, such as double-lock
standing seams, are not installed, copings should have
a continuous sheet membrane under the coping or the
roof flashing should be installed to run up and over the
parapet wall under the coping. These materials should
be capable of providing a secondary water barrier.
NRCA does not recommend caulking the joint between the bottom of the exterior vertical face of a coping cap or metal edge flashing and building exterior
because it does not add to the waterproofing integrity
of the assembly and it may, in fact, hinder the ability of
the wall construction to dissipate moisture. However,
the flashing must have sufficient lap (4 in. [100 mm]
min.) over the wall material to prevent water from being blown under the coping or flashing during wind
driven rain events.
There are three common methods for securing metal
copings, which can be seen in Fig. 5.8.5. These methods are described as follows:
Option A: A continuous cleat is fastened on both

5

Fig. 5.8.6 Alternate parapet and roofing details.
60 mil pvc flashing over precast panel joints,
secure with adhesive

Continuous metal
cap flashing

Joint
gasket

Sealant
1/8" thick continuous
termination bar

If parapet wall is too
high, use clamp bars
to hold plastic flashing
in place over panel
joints

the exterior and interior faces of the wall. The hem at
the drip edge of the coping metal on the interior side
should be fabricated slightly open for ease of application. After the exterior face of the coping is hooked
to the cleat, the interior face (roof side) is secured by
crimping the open hem of the coping to the cleat.
Option B: A continuous cleat is fastened to the exterior face of the wall, and clips are installed intermittently on the interior face (roof side) of the wall. After
the exterior face of the coping is hooked to the cleat,
the interior face (roof side) is secured by engaging the
clips to the bottom edge of the interior face of the
metal coping.
Option C: A continuous cleat is fastened on the exterior face of the wall. After the exterior face of the coping is hooked to the cleat, the interior face (roof side)
of the coping is secured to the parapet with gasketed
fasteners.
The waterproofing of parapet joints is important because they are exposed to weathering from all directions
and require regular checking for performance. Figure
5.8.6 shows a possible solution for joint and roofing
details involving a precast concrete parapet joint and
cap flashing. The metal cap flashing should have at
least a 4 in. (100 mm) exposed apron extending down

ARCHITECTURAL PRECAST CONCRETE

|

541

5

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.3 Parapet Details / 5.8.4 Scuppers

Fig. 5.8.7 Through-wall scupper.
Sealant
Sealant or
Base Flashing
Exterior
Closure
Flange

Scupper Liner

Fig. 5.8.8 Through-wall scupper with conductor head.

1" (25 mm) Min.

Overflow
Outlet

Scupper Liner
Conductor Head

Downspout

NOTES
— Conductor head to be a min. 1" (25 mm) below top of scupper liner.
— In areas where ice forms during winter conditions, the position of the
conductor head relative to the scupper liner would allow for drainage
above ice dam(s).

over the base flashing. In most cases the metal counterflashing is attached to the inside face of the metal
wall cap flashing with sheet metal screws. The choice
of flashing material and/or its treatment against corrosion should be based on preventing potential staining
of the precast concrete surface (see Section 3.6.3).
If a two-stage joint is used, the sealant should be continued up the back of the parapet, along the top, and
down the front to overlap the rain barrier. The top and
back of the parapet normally receive a field-molded
sealant to form a flush joint. In any solution using twostage joints, it is important that the airseal closely abuts
the roof seal to complete the building envelope.
If the panel-to-panel caulk joint of a precast concrete
wall system fails, water intrusion behind the flashing
will occur. Therefore, some designers extend the roof-

542

|

ARCHITECTURAL PRECAST CONCRETE

ing system up the back of the precast concrete and
over the top, covering the entire back of the panel including the panel-to-panel caulk joint. The roof termination is completed at the top with a metal, stone, or
precast concrete coping piece to hide the termination.
Architects should be careful to consider the type
and location of the precast concrete connections; frequently these connections can puncture the flashing.
Working with the precast concrete engineer and roofing contractor during the shop-drawing connection
phase helps avoid or minimize this problem. Also they
should ensure that all precast concrete back surfaces
above the roof line are smooth trowelled.
The intersection of the roof and the wall is a common site for discontinuities in the thermal insulation
and air-barrier systems. The key issue for controlling
air leakage is sealing the wall air barrier to the roofing membrane, and doing so in a manner that will accommodate the differential movement that generally
occurs at this junction. To control condensation at this
junction, the vapor retarder needs to be kept warm
by a continuous layer of thermal insulation. Continuity
of the thermal insulation and air barrier system also
serves to control heat loss and reduces the potential
for condensation occurring at this location.

5.8.4 Scuppers
Scuppers are typically installed at roof edges, either
through the perimeter parapet wall or within a raisededge design (Fig. 5.8.7 and 5.8.8). Scuppers can be
the primary or secondary source of drainage; scuppers
used as secondary drainage devices are called overflow
scuppers. Either type of scupper can discharge water
into a conductor head and downspout, be extended to
allow unrestricted discharge of water, or can discharge
water directly into a gutter. Typically, a through-wall
scupper discharges into a conductor head and not a
gutter, whereas an overflow scupper has unrestricted
discharge. The conductor head should be wider than
the scupper opening, and conductor heads should be
attached to the wall securely with the appropriate fasteners (Fig. 5.8.8).
Where no conductor head and downspout are used,
it is recommended that the precast concrete scupper liner be extended beyond the face of the wall.
Extending the scupper liner and adding a drip edge
will help prevent water from wicking back under the

OTHER ARCHITECTURAL DESIGN CONSIDERATIONS
5.8.4 Scuppers

Fig. 5.8.9(a) & (b) Scuppers.

(a)

5

scupper and into the wall or down the face of the wall.
When a conductor head and downspout are not used
it is important to consider the path of discharge and
the effects this may have on the face of the wall and
adjacent structures.
Good precast concrete scupper details are shown in
Fig. 5.8.9(a) and (b). The scupper should be placed
away from panel joints to keep water from running
into these joints. If a collection box is used, the rim of
the conductor head should be installed a minimum of 1
in. (25 mm) below the bottom opening of the scupper
to assure that tolerances allow positive drainage of the
roof under all conditions. Also, consideration should
be given to the clearance necessary to accommodate
tolerances of the roof slab depth, roof insulation, and
roofing placement to allow proper drainage.
If the scupper is located within the panel face, a minimum blockout dimension of 6 in. (152 mm) is necessary
in order to minimize the plugging of the opening.

(b)

ARCHITECTURAL PRECAST CONCRETE

|

543

6

Asp Avenue
Parking Facility
Norman, Oklahoma;
Architect: FrankfurtShort-Bruza Associates.

CHAPTER SI X
GUIDE SPECIFICATION

F OR A R CH I T E C T UR A L PREC A ST C O N CRETE
6.1 GENERAL
This chapter provides a basis for specifying in-plant
fabrication, including product design not shown on
contract drawings, and field erection of architectural
precast concrete. It does not address structural precast
concrete, coatings, or sealing the joints between units.

6.2 DRAWINGS AND SPECIFICATIONS
6.2.1 Drawings
The Architect’s or Engineer’s drawings should show
panel locations and necessary sections and dimensions
to define the size and shape of the architectural precast
concrete units, indicate the location and size of reveals,
bullnoses, and joints (both functional and aesthetic),
and illustrate details between panels and adjacent materials. When more than one type of panel material or
finish is used, indicate the extent and location of each
type on the drawings. The location and details of applied and embedded items should be shown on the
drawings. Plans should clearly differentiate between
architectural and structural precast concrete if both
are used on the same project. The details of corners
of the structure and interfacing with other materials
should be illustrated. The aesthetic requirements and
design loads should be identified, and load support
points and space allowed for connections should be
indicated. The Engineer of Record needs to be aware
of the magnitude and direction of all anticipated loads
to be transferred from the architectural precast concrete components to the building structural framing
and their points of application. These loads should be
addressed in the bid documents. It is especially critical
that the Engineer of Record make provisions for stiffeners and bracing required to transfer the architectural
precast concrete loads to the structural frame.
There should be no gaps between the specifications
and drawings nor should they overlap; the specifications and drawings should be complementary.

6.2.2 Specifications
The type and quality of the materials incorporated
into the units, the design compressive strength of the
concrete, the finishes, and the tolerances for fabrica-

tion and erection should be described. In the event
a performance specification is used appropriate data
should be included for the precaster to assess the
scope and quality of the precast concrete units to be
fabricated.
Specifiers should consider permitting variations in production methods, structural design, materials, connection and erection techniques to accommodate varying
plant practices. Specifying the results desired without
specifically defining the manufacturing procedures will
ensure the best competitive bidding. Required submittals should also include range-bracketing samples for
color and texture.
The availability, quantity, performance, cost and production considerations of each ingredient and finish of
architectural precast concrete can have a large impact
on a project’s schedule and budget. Therefore, they
should be determined and specified for each specific
project before the project specifications are released.
The time and expense required to develop samples and
select concretee colors and textures can be considerable
and should not be underestimated by the design team.
The specification section should include requirements
for connection components embedded in the precast
concrete, related loose connection hardware, and any
special devices for lifting or erection, if required. Items to
be specified in other sections include building frame support provisions required to support units, including portions of connectors attached to the structure, joint sealing and final cleaning, and protection of the architectural
precast concrete.

6.2.3 Coordination
The responsibility for supply of precast concrete support items to be placed on or in the structure in order
to receive the architectural precast concrete units depends on the type of structure and varies with local
practice. Clearly specify responsibility for supply and
installation of pre-erection hardware. If not supplied
by the precast concrete fabricator, list supplier and installation requirements in related trade sections.
The type and quantity of hardware items required to
be cast into precast concrete units for the use of other

ARCHITECTURAL PRECAST CONCRETE

|

545

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
Coordination / Performance Specifications

trades should be clearly specified. Specialty items should
be required to be detailed, and supplied to precaster in
a timely manner by the trade requiring them. Verify that
materials specified in the section on flashing are galvanically compatible with cast in reglets or counterflashing
receivers. Check that concrete coatings, adhesives, and
sealants specified in other sections are compatible with
each other and with the form release agents and surfaces
to which they are applied.
Items mentioned in the Guide Specification as supply and/or installation by others should be mentioned
in the specifications covering the specific trades. Such
items may include:
• Cost of additional inspection by an independent
testing laboratory, if required.
• Hardware for interfacing with other trades (window, door, flashing, and roofing items).
• Placing of precast concrete hardware cast into or
attached to the structure, including tolerances for
such placing.
• Joint treatment for joints between precast concrete
and other materials.
• Access to building and floors.
• Power and water supply.
• Cleaning.
• Water repellent coatings.
• Plant-installed facing materials such as natural stone
and clay products.

6.2.4 Guide Specification Development:
This Guide Specification developed by PCI, is based on
MASTERSPEC® Section 034500 “Precast Architectural
Concrete,” and is used by permission of ARCOM.
MASTERSPEC® is a product of the American Institute
of Architects (AIA) and is exclusively produced and
published by ARCOM on a licensed user basis. For
further information, call 800-424-5080, or visit www.
arcomnet.com.

6.3 TYPES OF SPECIFICATIONS
The most common form of an architectural precast
concrete specification is by performance. The principal
advantage of performance specifying over prescriptive
is that it allows precasters to combine economy and
optimum quality, utilizing established tooling and production techniques not envisioned by the architect or
specifier.

546

|

ARCHITECTURAL PRECAST CONCRETE

Performance specifications define the scope of work
by the results desired. For example, architectural precast concrete performance specifications will establish:
(1) drawings that govern the design and arrangement
of the various wall components; (2) quality of materials
and types of finishes; (3) the loads and forces the wall
panels are required to support; and (4) insulating and
permeability requirements. In other words, they cover
the aesthetic, functional, and structural requirements
and define all limiting factors.
Performance specifications can achieve good results
as long as the architect identifies the purpose to be
served. Performance specifications often include appropriate quality control safeguards such as pre-qualification of precasters, pre-bid approval of materials
and samples, careful review of shop drawings, and
architect’s approval of initial production units.
An alternative form of specifying is the prescriptive
method. Prescriptive specifications typically contain
inflexible and too stringent requirements that can adversely affect a project’s budget and delivery schedule.
An example of prescriptive specifying would be preengineered cladding systems. In this example an owner
will engage a design firm to engineer a cladding system in order to shorten the time period necessary to
design and develop project shop drawings.
Performance specifications may create additional work
for the architect at the design stage, because the end
result must be clearly defined and frequently multiple
bid proposals must be assessed. The accepted proposal
will eventually become the standard for manufacturing.
However, this additional work in the early stages is generally offset by time saved later in detailing in the architect’s
office.
Performance specifications should define the scope
(statement of needs) and quality of the precast concrete at an early project stage. With performance specifications, the manufacturer is responsible for selecting
means and methods to achieve satisfactory results.
Properly prepared performance specifications should
conform to the following criteria:
1. They should clearly state all limiting factors such as
minimum or maximum thickness, depth, weight, tolerances, and any other limiting dimensions. Acceptable limits for
requirements not detailed should be clearly provided. These
limits may cover insulation (thermal and acoustical), interaction with other materials, services, and appearance.

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
Performance Specifications / 1.2 Summary

2. They should be written so that the scope is clearly
defined. Items both included and not included under
the scope of the precast concrete work must be identified and cross-referenced in the project documents.
3. The architect should request samples, design and detail
submissions from prospective bidders, and make pre-bid
approval of such submissions a prerequisite for bidding.
4. If such requests for pre-bid approvals form a part
of the specifications, the architect should adhere to the
following:

6

a. Sufficient time must be allowed for the precaster to prepare and submit samples or information
for approval by the architect. Approval should be
conveyed to the manufacturer in writing with sufficient time to allow completion of an estimate and
submittal of a bid.
b. All proprietary pre-bid submittals should be treated
in confidence and the individual precaster’s original solutions or techniques protected both before
and after bidding.

6.4 Guide Specification
This Guide Specification is intended to be used as a basis for the development of an office master specification or
in the preparation of performance specifications for a particular project. In either case, this Guide Specification
must be edited to fit the conditions of use. Particular attention should be given to the deletion of inapplicable
provisions or inclusion of additional appropriate requirements. Coordinate the specifications with the information
shown on the Contract Drawings to avoid duplication or conflicts.
Shaded portions are Notes to the Specification Writer.

SECTION 034500
PRECAST ARCHITECTURAL CONCRETE
This Section uses the term “Architect.” Change this term to match that used to identify the design professional
as defined in the General and Supplementary Conditions of the contract. Verify that Section titles referenced in
this Section are correct for this Project’s Specifications; Section titles may have changed.

PART 1 – GENERAL
1.1 R
 ELATED DOCUMENTS
A.	Drawings and general provisions of the Contract, including General and Supplementary Conditions and
Division 01 Specification Sections, apply to this Section.

1.2 S UMMARY
A.	This section includes the performance criteria, materials, production, and erection of architectural precast concrete for the entire project. The work performed under this Section includes all labor, material,
equipment, related services, and supervision required for the manufacture and erection of the architectural precast concrete work shown on the Contract Drawings.
Adjust list below to suit Project. Delete paragraph below if not listing type of units.
B.
		
		
		
		

This Section includes the following:
1. Architectural precast concrete cladding (and loadbearing) units.
2. Insulated, architectural precast concrete units.
3. Clay product-faced, architectural precast concrete units.
4. Stone veneer-faced, architectural precast concrete units.

ARCHITECTURAL PRECAST CONCRETE

|

547

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.2 Summary / 1.4 Performance Requirements

C.

Related Sections include the following:

List below only products and construction that the reader might expect to find in this Section but are specified elsewhere. Other sections of the specifications not referenced below, also apply to the extent required for
proper performance of this work.
		
		
		
		
		
		
		
		
		
		
		
		
		

1. Division 03 Section “Cast-in-Place Concrete” for installing connection anchors in concrete.
2. Division 03 Section “Glass-Fiber-Reinforced Concrete (GFRC).”
3. Division 04 Section “Exterior Stone Cladding” for furnishing stone facings and anchorages.
4. Division 04 Section “Cast Stone Masonry” for wet or dry cast stone facings, trim, and accessories.
5.	Division 04 Section “Unit Masonry Assemblies” for full-thickness brick facing, mortar, inserts, and
anchorages.
6.	Division 05 Section “Structural Steel Framing” for furnishing and installing connections attached to
structural-steel framing.
7.	Division 05 Section “Metal Fabrications” for furnishing and installing loose hardware items, kickers,
and other miscellaneous steel shapes.
8. Division 07 Section “Water Repellents” for water-repellent finish treatments.
9. Division 07 Section “Sheet Metal Flashing and Trim” for flashing receivers and reglets.
10. Division 07 Section “Joint Sealants” for elastomeric joint sealants and sealant backings.
11.	Division 08 Section “Aluminum Windows” for windows set into architectural precast concrete units.
12.	Division 09 Section “Tiling” for ceramic tile setting materials and installation.
13.	Division 11 Section “Window Washing Equipment” for tie-backs located in architectural precast
concrete units.

1.3 D
 EFINITION
Retain paragraph below if a design reference sample has been preapproved by Architect and is available for
review.
A.	Design Reference Sample: Sample of approved architectural precast concrete color, finish and texture,
preapproved by Architect.

1.4 P
 ERFORMANCE REQUIREMENTS
Retain this Article if delegating design responsibility for architectural precast concrete units to Contractor. AIA
Document A201 requires Owner or Architect to specify performance and design criteria.
A.	Structural Performance: Provide architectural precast concrete units and connections capable of withstanding the following design loads within limits and under conditions indicated:
		
1. Loads: As indicated.
Retain paragraph above if design loads are shown on Drawings; delete subparagraph above and retain paragraph
and applicable subparagraphs below if including design loads here. Revise requirements below to suit Project,
and add other performance and design criteria if applicable.
B.	Structural Performance: Provide architectural precast concrete units and connections capable of withstanding the following design loads within limits and under conditions indicated:

548

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.4 Performance Requirements

6

As a minimum dead loads include panel weight and the weight(s) of the materials that bear on them.
		
		
		

		

1. Dead Loads: 
2. Live Loads: 
3.	Wind Loads: 
4.	Seismic Loads: 

Precast specific loads may include blast loads.
		
		

5.	Project Specific Loads: 
6.	Design precast concrete units and connections to maintain clearances at openings, to allow for
fabrication and construction tolerances, to accommodate live-load deflection, shrinkage and creep
of primary building structure, and other building movements as follows:

Indicate locations here or on Drawings if different element structural, shrinkage, creep or thermal movements are
anticipated for different building elements. If preferred, change deflection limits in subparagraph below to ratios
such as L/300 for floors and L/200 for roofs. Verify all building frame movements with the Engineer of Record.
			
			
			

a. Upward and downward movement of (1/2 in. [13 mm]) (3/4 in. [19 mm]) (1 in. [25 mm]).
b.	Overall building drift: 
c. Interstory building drift: 

Temperature value in first subparagraph below is suitable for most of the U.S. based on assumed design nominal
temperature of 70 °F (21 °C). Revise subparagraph below to suit local conditions. Temperature data are available
from National Oceanic and Atmospheric Administration at www.ncdc.noaa.gov.
		

7.	Thermal Movements: Provide for in-plane thermal movements resulting from annual ambient temperature changes of (80 °F [26 °C]) .

Delete subparagraph below if fire resistance rating is not required. Fire ratings depend on occupancy and
building construction type, and are generally a building code requirement. When required, fire-rated products
should be clearly identified on the design drawings.
		

8.	Fire Resistance Rating: Select material and minimum thicknesses to provide (1)(2) 
- hour fire rating.

Delete subparagraph below if window washing system is not required. Indicate design criteria here or on
Drawings for window washing system, including material and equipment..
		

9.	Window Washing System: Design precast concrete units supporting window washing system indicated to resist pull-out and horizontal shear forces transmitted from window washing equipment.

Retain subparagraph below if stone veneer–faced precast concrete units are used on project.
		

10.	Stone to Precast Concrete Anchorages: Provide anchors, as determined through Owner’s or stone
supplier testing, in numbers, types, and locations required to satisfy specified performance criteria.

ARCHITECTURAL PRECAST CONCRETE

|

549

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.4 Performance Requirements / 1.5 Submittals

Delete subparagraph below if precast concrete units are not used in parking structure to resist impact load.
Local codes may have requirements that vary from those listed.
		

11.	Vehicular Impact Loads: Design spandrel beams acting as vehicular barriers for passenger cars to resist a
single load of (6,000 lb [26.7 kN])  service load and (10,000 lb [44.5 kN]) 
ultimate load applied horizontally in any direction to the spandrel beam, with anchorages or attachments
capable of transferring this load to the structure. Design spandrel beams, assuming the load to act at a
height of 18 in. (457 mm) above the floor or ramp surface on an area not to exceed 1 ft2 (0.09 m2).

1.5 S UBMITTALS
A.	Product Data: For each type of product indicated. Retain quality control records and certificates of compliance for 5 years after completion of structure.
B.	LEED Submittals:
Retain subparagraph below if recycled content is required for LEED-NC or LEED-CI Credits MR 4.1 and MR 4.2.
An alternative method of complying with Credit MR 4.1 and MR 4.2 requirements is to retain requirement in
Division 01 SECTION “Sustainable Design Requirements” that gives Contractor the option and responsibility for
determining how Credit MR 4.1 and MR 4.2 requirements will be met.
1.	Product Data for Credit MR 4.1[and Credit MR 4.2]: For products having recycled content, documentation
indicating percentages by weight of postconsumer and preconsumer (post-industrial) recycled content per
unit of product.
			
a.	Indicate recycled content; indicate percentage of pre-consumer and post-consumer recycled content
per unit of product.
			
b.	Indicate relative dollar value of recycled content product to total dollar value of product included in
project.
			
c.	If recycled content product is part of an assembly, indicate the percentage of recycled content product
in the assembly by weight.
			
d.	If recycled content product is part of an assembly, indicate relative dollar value of recycled content
product to total dollar value of assembly.
		
2.	Product Data for Credit MR 5.1 [and Credit MR 5.2]: For local and regional material extracted/harvested
and manufactured within a 500 mile radius from the project site.
			
a.	Indicate location of extraction, harvesting, and recovery; indicate distance between extraction, harvesting, and recovery and the project site.
			
b.	Indicate location of manufacturing facility; indicate distance between manufacturing facility and the
project site.
			
c.	Indicate dollar value of product containing local/regional materials; include materials cost only.
			
d.	Where product components are sourced or manufactured in separate locations, provide location
information for each component. Indicate the percentage by weight of each component per unit of
product.
		

Retain subparagraph below if environmental data is required in accordance with Table 1 of ASTM E 2129. Concrete is
relatively inert once cured. Admixtures, form release agents, and sealers may emit VOCs, especially during the curing
process; however, virtually all emissions are eliminated before enclosing the building.
		

550

|

3.	Include MSDS product information showing that materials meet any environmental performance goals such
as biobased content.

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.5 Submittals

		
		

6

4.	For projects using FSC certified formwork, include chain-of-custody documentation with certification numbers for all certified wood products.
5.	For projects using reusable formwork, include data showing how formwork is reused.

C.	Design Mixtures: For each precast concrete mixture. Include results of compressive strength and waterabsorption tests.
D.	Shop (Erection) Drawings: Detail fabrication and installation of architectural precast concrete units. Indicate locations, plans, elevations, dimensions, shapes, and cross-sections of each unit. Indicate aesthetic intent including
joints, rustications or reveals, and extent and location of each surface finish. Indicate details at building corners.
Delete subparagraphs below not applicable to Project.
		
		
		
		
		
		
		
		
		
		
		

1.	Indicate separate face and backup mixture locations and thicknesses.
2.	Indicate welded connections by AWS standard symbols and show size, length, and type of each
weld. Detail loose and cast-in hardware and connections.
3.	Indicate locations, tolerances, and details of anchorage devices to be embedded in or attached to
structure or other construction.
4.	Indicate locations, extent, and treatment of dry joints if two-stage casting is proposed.
5.	Indicate plans and/or elevations showing unit location and dimensions, erection sequences, and bracing plan
for special conditions.
6.	Indicate location of each architectural precast concrete unit by same identification mark placed on unit.
7.	Indicate relationship of architectural precast concrete units to adjacent materials.
8.	Indicate locations and details of clay product units, including corner units and special shapes with
dimensions, and joint treatment.
9.	Indicate locations and details of stone veneer-facings, stone anchors, and joint widths.
10.	Coordinate and indicate openings and inserts required by other trades.
11.	Design Modifications: If design modifications are proposed to meet performance requirements and
field conditions, notify the Architect and submit design calculations and Shop Drawings. Do not
adversely affect the appearance, durability, or strength of units when modifying details or materials
and maintain the general design concept.

Retain subparagraph below if retaining “Performance Requirements” Article. Delete or modify if Architect assumes
or is required by law to assume design responsibility.
		

12.	Comprehensive engineering design (signed and sealed) (certified) by qualified professional
engineer responsible for its preparation licensed in the jurisdiction in which the project is located.
Show governing panel types, connections, and types of reinforcement, including special reinforcement
such as epoxy coated carbon fiber grid. Indicate location, type, magnitude, and direction of all loads
imposed on the building structural frame by the architectural precast concrete.

Retain paragraph and subparagraphs below if finishes, colors, and textures are preselected, specified, or scheduled. Coordinate with sample panels and range samples in “Quality Assurance” Article.
E.	Samples: Design reference samples for initial verification of design intent, approximately 12 x 12 x 2 in.
(300 x 300 x 50 mm), representative of finishes, color, and textures of exposed surfaces of architectural
precast concrete units.
		
1.	When back face of precast concrete unit is to be exposed, include Samples illustrating workmanship,
color, and texture of the backup concrete as well as facing concrete.

ARCHITECTURAL PRECAST CONCRETE

|

551

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.5 Submittals / 1.6 Quality Assurance

Retain subparagraph below if Samples of thin brick facings are required.
		

2.	Samples for each brick unit required, showing full range of color and texture expected. Include
Sample showing color, geometry, and texture of joint treatment.

Retain paragraph below if procedures for welder certification are retained in “Quality Assurance” Article.
F.	Welding Certificates: Copies of certificates for welding procedure specifications (WPS) and personnel certification.
Manufacturer should have a minimum of 2 years of production experience in architectural precast concrete
work comparable to that shown and specified, in not less than three projects of similar scope with the Owner
or Architect determining the suitability of the experience.
G.	Qualification Data: For firms and persons specified in “Quality Assurance” Article to demonstrate their
capabilities and experience. Include list of completed projects with project names and addresses, names
and addresses of architects and owners, and other information specified.
Delete test reports paragraph below if not required.
H.	Material Test Reports: From an accredited testing agency, indicating and interpreting test results of the
following, for compliance with requirements indicated:
Retain paragraph above or below.
I.	Material Certificates. For the following items signed by manufacturers:
Retain list below with either paragraph above. Edit to suit Project.
		
		
		
		
		
		
		
		

1.
2.
3.
4.
5.
6.
7.
8.

Cementitious materials.
Reinforcing materials including prestressing tendons.
Admixtures.
Bearing pads.
Structural-steel shapes and hollow structural steel sections.
Insulation
Clay product units and accessories.
Stone anchors.

Retain paragraph below if Contractor is responsible for field quality-control testing. Retain option if Contractor
is responsible for special inspections.
J.	Field quality-control test [and special inspections] reports.

1.6 Q
 UALITY ASSURANCE
Erector should have a minimum of 2 years of experience in architectural precast concrete work comparable to
that shown and specified in not less than three projects of similar scope with the Owner or Architect determining the suitability of the experience. The inclusion of erection in the precast concrete contract should be governed by local practices. Visit the PCI website at www.pci.org for current listing of PCI- Qualified and Certified
Erectors. Retain first paragraph below if PCI-Certified Erector is not available in project location.

552

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.6 Quality Assurance

6

A.	Erector Qualification: A precast concrete erector with all erecting crews Qualified and designated, prior to beginning work at project site, by PCI’s Certificate of Compliance to erect (Category A [Architectural Systems] for
non-load-bearing members) (Category S2 [Complex Structural Systems] for load-bearing members).
B.	Erector Certification: A precast concrete erector with erecting organization and all erecting crews
Certified and designated, prior to beginning work at project site, by PCI’s Certificate of Compliance
to erect (Category A [Architectural Systems] for non-load-bearing members) (Category S2
[Complex Structural Systems] for load-bearing members).
Retain first paragraph below if PCI-Qualified or Certified Erector is not available in Project location. Basis of audit is PCI MNL-127, Erector’s Manual – Standards and Guidelines for the Erection of Precast Concrete Products.
C.	Erector Qualifications: A precast concrete erector who has retained a “PCI-Certified Field Auditor”, at
erector’s expense, to conduct a field audit of a project in the same category as this Project prior to start
of erection and who can produce an Erector’s Post Audit Declaration.
D.	Fabricator Qualifications: A firm that complies with the following requirements and is experienced in
producing architectural precast concrete units similar to those indicated for this Project and with a record
of successful in-service performance.
		
1.	Assumes responsibility for engineering architectural precast concrete units to comply with performance requirements. This responsibility includes preparation of Shop Drawings and comprehensive
engineering analysis by a qualified professional engineer.
Delete subparagraph above and below if Precaster is not required to engage the services of a qualified
professional engineer and if submission of a comprehensive engineering analysis is not retained in
“Submittals” Article.
		

		
		

2.	Professional Engineer Qualifications: A professional engineer who is legally qualified to practice in the
jurisdiction where Project is located and who is experienced in providing engineering services of the
kind indicated. Engineering services are defined as those performed for installations of architectural
precast concrete that are similar to those indicated for this Project in material, design, and extent.
3.	Participates in PCI’s Plant Certification program (at the time of bidding) and is designated a PCICertified plant for Group A, Category A1- Architectural Cladding and Loadbearing Units.
4. Has sufficient production capacity to produce required units without delaying the Work.

Delete subparagraph below if fabricators are not required to be registered with and approved by authorities
having jurisdiction. List approved fabricators in Part 2 if required.
		

5.

Is registered with and approved by authorities having jurisdiction.

Retain first paragraph below if quality assurance testing in addition to that provided by the PCI Certification
Program is required. Testing agency if required, is normally engaged by Owner.
E.	Testing Agency Qualifications: An independent testing agency (acceptable to authorities having
jurisdiction), qualified according to ASTM C 1077 and ASTM E 329 to conduct the testing indicated.
F.	Design Standards: Comply with ACI 318 (ACI 318M) and design recommendations of PCI MNL 120,
PCI Design Handbook – Precast and Prestressed Concrete, applicable to types of architectural precast
concrete units indicated.

ARCHITECTURAL PRECAST CONCRETE

|

553

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.6 Quality Assurance

G.	Quality-Control Standard: For manufacturing procedures and testing requirements, quality-control
recommendations, and dimensional tolerances for types of units required, comply with PCI MNL 117,
Manual for Quality Control for Plants and Production of Architectural Precast Concrete Products.
Delete paragraph below if no welding is required. Retain “Welding Certificates” paragraph in “Submittals”
Article if retaining below. AWS states that welding qualifications remain in effect indefinitely unless welding
personnel have not welded for more than six months or there is a specific reason to question their ability.
H.	Welding: Qualify procedures and personnel according to AWS D1.1/D1.1M, “Structural Welding Code
– Steel”; and AWS D1.4, “Structural Welding Code – Reinforcing Steel.”
Retain paragraph below if fire-rated units or assemblies are required. Select either PCI MNL 124 or
ACI 216.1/TMS 0216.1 or retain both if acceptable to authorities having jurisdiction.
I.	Fire Resistance: Where indicated, provide architectural precast concrete units whose fire resistance meets
the prescriptive requirements of the governing code or has been calculated according to (PCI MNL 124,
Design for Fire Resistance of Precast Prestressed Concrete) (ACI 216.1/TMS 0216.1, Standard
Method for Determining Fire Resistance of Concrete and Masonry Construction Assemblies)
and is acceptable to authorities having jurisdiction.

PCI recommends review of preproduction sample panels or first production unit. Revise size and number of
sample panels in paragraph below to suit Project.
J.	Sample Panels: After sample approval and before fabricating architectural precast concrete units, produce a minimum of (two)  sample panels approximately (16 ft2 [1.5 m2])  in area for review by Architect. Incorporate full-scale details of architectural features, finishes,
textures, and transitions in the sample panels.
		
1. Locate panels where indicated or, if not indicated, as directed by Architect.
		
2. 	Damage part of an exposed-face surface for each finish, color, and texture, and demonstrate adequacy of repair techniques proposed for repair of surface blemishes.
		
3.	After acceptance of repair technique, maintain one sample panel at the manufacturer’s plant and
one at the Project site in an undisturbed condition as a standard for judging the completed Work.
		
4. Demolish and remove sample panels when directed.
PCI recommends production of finish and texture range samples when color and texture uniformity concerns
could be an issue, Architect or precaster has not had previous experience with the specified mixture and finish,
or a large project has multiple approving authorities.
K.	Range Sample Panels: After sample panel approval and before fabricating architectural precast concrete units, produce a minimum of (three)(five) samples, approximately (16 ft2 [1.5
m2])  in area, representing anticipated range of color and texture on Project’s units.
Maintain samples at the manufacturer’s plant as color and texture acceptability reference.
Delete paragraph and subparagraphs below if sample panels and range samples above will suffice and added
expense of mockups is not required. If retaining, indicate location, size, and other details of mockups on
Drawings or by inserts. Revise wording if only one mockup is required.

554

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.6 Quality Assurance

6

L.	Mockups: After sample panel (and range sample) approval but before production of architectural
precast concrete units, construct full-sized mockups to verify selections made under sample submittals
and to demonstrate aesthetic effects and set quality standards for materials and execution. Mockups to
be representative of the finished work in all respects including (glass, aluminum framing, sealants)
 and architectural precast concrete complete with anchors, connections,
flashings, and joint fillers as accepted on the final Shop Drawings. Build mockups to comply with the
following requirements, using materials indicated for the completed work:
Revise or delete subparagraphs below to suit Project.
1.	Build mockups in the location and of the size indicated or, if not indicated, as directed by Architect.
2. Notify Architect in advance of dates and times when mockups will be constructed.
3.	Obtain Architect’s approval of mockups before starting production fabrication of precast concrete units.
4.	Maintain mockups during construction in an undisturbed condition as a standard for judging the
completed Work.
5. Demolish and remove mockups when directed.

		
		
		
		
		

Retain first subparagraph below if mockups are erected as part of building rather than separately and the
intention is to make an exception to the default requirement in Division 01 Section “Quality Requirements”
for demolishing and removing mockups when directed, unless otherwise indicated.
		
		

6.	Approved mockups may become part of the completed Work if undamaged at the time of
Substantial Completion.
7.	Approval of mockups does not constitute approval of deviations from the Contact Documents
unless such deviations are specifically approved by Architect in writing.

Delete paragraph below if mockup above is to be used for Testing Mockup or if testing is not required. If
retaining paragraph and subparagraphs below, determine where preconstruction testing will be specified and
include requirements in that Section. Requirements in paragraph below are limited to building a preconstruction testing mockup at a testing agency’s facility.
M.	Preconstruction Testing Mockup: Provide a full-size mockup of architectural precast concrete indicated
on Drawings for preconstruction testing. Refer to Division [01][08]  Section
“” for preconstruction testing requirements.
Revise or delete subparagraphs below to suit Project. Coordinate with other Sections that include construction
to be included in a preconstruction testing mockup to clearly indicate extent of work required in this Section
		

		

1.	Build preconstruction testing mockup as indicated on Drawings including [glass, aluminum framing, sealants,]  and architectural precast concrete complete with anchors,
connections, flashings, and joint fillers.
2.	Build preconstruction testing mockup at testing agency facility.

Delete paragraph below if Work of this Section is not extensive or complex enough to justify a preinstallation
conference. If retaining, coordinate with Division 01.
N.	Preinstallation Conference: Conduct conference at Project site to comply with requirements in
Division 01 Section “Project Management and Coordination.”

ARCHITECTURAL PRECAST CONCRETE

|

555

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
1.7 Product Delivery, Storage, And Handling / 2.1 Fabricators

1.7 P
 RODUCT DELIVERY, STORAGE, AND HANDLING
A.	Store units with adequate dunnage and bracing, and protect units to prevent contact with soil, to
prevent staining, and to prevent cracking, distortion, warping, or other physical damage.
B.

Place stored units so identification marks are clearly visible, and units can be inspected.

C.	Deliver architectural precast concrete units in such quantities and at such times to ensure compliance
with the agreed project schedule and proper setting sequence and also to limit unloading units temporarily on the ground or other rehandling.
D.

Support units during shipment on non-staining shock- absorbing material.

E.	Handle and transport units in a position consistent with their shape and design in order to avoid
excessive stresses that could cause cracking or damage.
F.

Lift and support units only at designated points indicated on Shop Drawings.

1.8 S EQUENCING
Coordination and responsibility for supply of items to be placed on or in the structure to allow placement of precast
concrete units depends on type of structure and varies with local practice. Clearly specify responsibility for supply and
installation of hardware. If not supplied by precaster, supplier should be listed and requirements included in related
trade sections. Ensure that type and quantity of hardware items to be cast into precast concrete units for use by other
trades are specified or detailed in Contract Drawings and furnished to precaster, with instructions, in a timely manner
in order not to delay the Work.
A.	Furnish loose connection hardware and anchorage items to be embedded in or attached to other construction without delaying the Work. Provide locations, setting diagrams, templates, instructions, and
directions, as required, for installation.

PART 2 – PRODUCTS
2.1 F ABRICATORS
Delete this Article unless naming fabricators. See PCI’s magazine ASCENT or visit PCI’s website at www.pci.org
for current PCI-Certified plant listings.
A.	Available Fabricators: Subject to compliance with requirements, fabricators offering products that may
be incorporated into the Work include, but are not limited to, the following:
Retain above for nonproprietary or below for semiproprietary specification. If above is retained, include
procedure for approval of other fabricators in Instructions to Bidders. See Division 01 Section “Product
Requirements.”
B.	Fabricators: Subject to compliance with requirements, provide products by one of the following:
		
1.	

556

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.2 Mold Materials / 2.3 Reinforcing Materials

6

2.2 M
 OLD MATERIALS
A.	Molds: Rigid, dimensionally stable, non-absorptive material, warp and buckle free, that will provide
continuous and true precast concrete surfaces within fabrication tolerances indicated; nonreactive with
concrete and suitable for producing required finishes.
		
1.	Form-Release Agent: Commercially produced form-release agent that will not bond with, stain, or
adversely affect precast concrete surfaces and will not impair subsequent surface or joint treatments
of precast concrete.
Delete paragraph below if not using form liners. Form liners may be used to achieve a special off-the-form finish
or to act as a template for thin or half-brick facings. Revise to add description of selected form liner, if required.
B.	Form Liners: Units of face design, texture, arrangement, and configuration (indicated) (to match those
used for precast concrete design reference sample). Provide solid backing and form supports to
ensure that form liners remain in place during concrete placement. Use manufacturer’s recommended
form-release agent that will not bond with, stain, or adversely affect precast concrete surfaces and will
not impair subsequent surface or joint treatments of precast concrete.

Retain paragraph below if surface retarder is applied to molds to help obtain exposed aggregate finish.
C.	Surface Retarder: Chemical set retarder, capable of temporarily delaying setting of newly placed concrete
to depth of reveal specified.

2.3 R
 EINFORCING MATERIALS
Retain first paragraph below if recycled content is required for LEED-NC or LEED-CI Credits MR 4.1 and MR 4.2.
USGBC allows a default value of 25 percent to be used for steel, without documentation; higher percentages can be
claimed if they are supported by appropriate documentation. The Steel Recycling Institute indicates that reinforcing
bars are made by the electric arc furnace method, which typically has 67 percent post-consumer recycled content
and 6.5 percent pre-consumer recycled content.
A.	Recycled Content of Steel Products: Provide products with an average recycled content of steel products so postconsumer recycled content plus one-half of preconsumer recycled content is not less than
[25][60]  percent.
Select one or more of the paragraphs in this Article to suit steel reinforcement requirements. If retaining Part 1
“Performance Requirements” Article, consider reviewing selections with fabricators.
B.

Reinforcing Bars: ASTM A 615/A 615M, Grade 60 (Grade 420), deformed.

Retain paragraph below for reinforcement that is welded or if added ductility is sought.
C. Low-Alloy-Steel Reinforcing Bars: ASTM A 706/A 706M, deformed.
Retain galvanized reinforcement in paragraph below where corrosive environment or severe exposure conditions justify extra cost. The presence of chromate film on the surface of the galvanized coating is usually visible
as a light yellow tint on the surface. ASTM B 201 describes a test method for determining the presence of
chromate coatings.

ARCHITECTURAL PRECAST CONCRETE

|

557

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.3 Reinforcing Materials / 2.5 Concrete Materials

D.	Galvanized Reinforcing Bars: (ASTM A 615/A 615M, Grade 60 [Grade 420]) (ASTM A 706/A 706M),
deformed bars, ASTM A 767/A 767M Class II zinc-coated, hot-dip galvanized and chromate wash
treated after fabrication and bending.
Consider using epoxy-coated reinforcement where corrosive environment or severe exposure conditions justify
extra cost. In first paragraph below, retain ASTM A 775/A 775M for a bendable epoxy coating; retain ASTM A
934/A 934M for a nonbendable epoxy coating.
E.	Epoxy-Coated Reinforcing Bars: (ASTM A 615/A 615M, Grade 60 [Grade 420]) (ASTM A 706/A
706M), deformed bars, (ASTM A 775/A 775M) or (ASTM A 934/A 934M) epoxy coated.
F.	Steel Bar Mats: ASTM A 184/A 184M, fabricated from (ASTM A 615/A 615M, Grade 60 [Grade 420])
(ASTM A 706/A 706M) deformed bars, assembled with clips.
Select one or more of the paragraphs below to suit steel reinforcement requirements. If retaining Part 1
“Performance Requirements” Article, consider reviewing selections with fabricators.

G.	Plain-Steel Welded Wire Reinforcement: ASTM A 185, fabricated from (as-drawn) (galvanized and
chromate wash treated) steel wire into flat sheets.
H.	Deformed Steel Welded Wire Reinforcement: ASTM A 497/A 497M, flat sheet.
I.	Epoxy Coated-Steel Welded Wire Reinforcement: ASTM A 884/A 884M Class A coated, (plain)
(deformed), flat sheet, Type (1 bendable) (2 non-bendable) coating.
J.	Supports: Suspend reinforcement from back of mold or use bolsters, chairs, spacers, and other devices for spacing, supporting, and fastening reinforcing bars and welded wire reinforcement in place according to PCI MNL 117.

2.4 P
 RESTRESSING TENDONS
Retain this Article if precast concrete units will be prestressed, either pretensioned or post-tensioned. ASTM A
416/A 416M establishes low-relaxation strand as the standard.
A.	Prestressing Strand: ASTM A 416/A 416M, Grade 270 (Grade 1860), uncoated, 7-wire, low-relaxation strand.
B.	Unbonded Post-Tensioning Strand: ASTM A 416/A 416M with corrosion inhibitor coating conforming
to ASTM D1743, Grade 270 (Grade 1860), 7-wire, low-relaxation strand with polypropylene tendon
sheathing. Include anchorage devices.
C.	Post-Tensioning Bars: ASTM A 722, uncoated high strength steel bar.

2.5 C ONCRETE MATERIALS
Delete materials in this Arcticle that are not required; revise to suit Project.
A.

Portland Cement: ASTM C150, Type I or III.

Select portland cement color from options in subparagraph below. Mixing with white cement will improve
color uniformity of gray cement. White cement has greater color consistency than gray cement and should be
used for pastel colors. For darker colors, the variations of gray cement have less effect on the final color hue.

558

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.5 Concrete Materials

		

6

1.	For surfaces exposed to view in finished structure, use (gray) (or) (white), of same type, brand,
and mill source throughout the precast concrete production.

Delete subparagraphs below if only gray cement is selected in paragraph above. Retain below if face mixture
uses white cement but gray cement will be permitted in backup mixture.
		
B.

2.

Standard gray portland cement may be used for non-exposed backup concrete.

Supplementary Cementitious Materials.

Prior to selecting mineral or cementitious materials from four subparagraphs below consult local precasters. These
materials may affect concrete appearance, set times and cost. Where appearance is an important factor, it is recommended that fly ash and gray silica fume not be permitted for exposed exterior surfaces. White silica fume is available.
		
		
		
		

1.
2.
3.
4.

Fly Ash: ASTM C 618, Class C or F with maximum loss on ignition of 3%.
Metakaolin: ASTM C 618, Class N.
Silica Fume: ASTM C 1240 with optional chemical and physical requirements.
Ground Granulated Blast-Furnace Slag: ASTM C 989, Grade 100 or 120.

ASTM C 33 limits deleterious substances in coarse aggregate depending on climate severity and in-service location of concrete. Class 5S is the most restrictive designation for architectural concrete exposed to severe weathering. PCI MNL 117 establishes stricter limits on deleterious substances for fine and coarse aggregates.
C.	Normalweight Aggregates: Except as modified by PCI MNL 117, ASTM C 33, with coarse aggregates
complying with Class 5S. Provide and stockpile fine and coarse aggregates for each type of exposed finish from a single source (pit or quarry) for Project.
Revise subparagraph below and add descriptions of selected coarse- and fine-face aggregate colors, sizes, and
sources if required.
		

1.	Face-Mixture Coarse Aggregates: Selected, hard, and durable; free of material that reacts with cement or causes staining; to match selected finish sample.

Retain one option from first subparagraph below or insert gradation and maximum aggregate size if known.
Fine and coarse aggregates are not always from same source.
			
		

a.	Gradation: (Uniformly graded) (Gap graded) (To match design reference sample).

2.	Face-Mixture Fine Aggregates: Selected, natural, or manufactured sand of a material compatible
with coarse aggregate to match selected Sample finish.

Delete subparagraph below when architectural requirements dictate that face-mixture be used throughout.
		

3.

Backup Concrete Aggregates: ASTM C 33 or C 330.

Lightweight aggregates in a face-mixture are not recommended in cold or humid climates (if exposed to the
weather) unless their performance has been verified by tests or records of previous satisfactory usage in similar
environments. If normalweight aggregates are used in face-mixture, lightweight aggregates in the backup
mixture are not recommended due to panel bowing potential.

ARCHITECTURAL PRECAST CONCRETE

|

559

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.5 Concrete Materials / 2.6 Steel Connection Materials

D.	Lightweight Aggregates: Except as modified by PCI MNL 117, ASTM C 330 with absorption less than 11%.
Delete first paragraph below if coloring admixture is not required. Add color selection if known.
E.	Coloring Admixture: ASTM C 979, synthetic or natural mineral–oxide pigments or colored water-reducing
admixtures, temperature stable, and non-fading.
F.	Water: Potable; free from deleterious material that may affect color stability, setting, or strength of concrete
and complying with ASTM C 1602/C 1602M and chemical limits of PCI MNL 117.
Delete paragraph below if air entrainment is not required. Air entrainment should be required to increase resistance to freezing and thawing where environmental conditions dictate.
G.	Air-Entraining Admixture: ASTM C 260, certified by manufacturer to be compatible with other required
admixtures.
H.	Chemical Admixtures: Certified by manufacturer to be compatible with other admixtures and to not
contain calcium chloride, or more than 0.15% chloride ions or other salts by weight of admixture.
Limit chemical admixture types if required.
		
		
		
		
		
		
		

1.
2.
3.
4.
5.
6.
7.

Water-Reducing Admixture: ASTM C 494/C 494M, Type A.
Retarding Admixture: ASTM C 494/C 494M, Type B.
Water-Reducing and Retarding Admixture: ASTM C 494/C 494M, Type D.
Water-Reducing and Accelerating Admixture: ASTM C 494/C 494M, Type E.
High-Range, Water-Reducing Admixture: ASTM C 494/C 494M, Type F.
High-Range, Water-Reducing and Retarding Admixture: ASTM C 494/C 494M, Type G.
Plasticizing Admixture for Flowable Concrete: ASTM C 1017/C 1017M.

2.6 S TEEL CONNECTION MATERIALS
Edit this Article to suit Project. Add other materials as required.
A.

Carbon-Steel Shapes and Plates: ASTM A 36/A 36M.

B.	Carbon-Steel Headed Studs: ASTM A 108, Grades 1010 through 1020, cold finished, AWS D1.1/ D1.1 M,
Type A or B, with arc shields and with minimum mechanical properties of PCI MNL 117, Table 3.2.3.

560

|

C.

Carbon-Steel Plate: ASTM A 283/A 283M.

D.

Malleable Iron Castings: ASTM A 47/A 47M, Grade 32510 or 35028.

E.

Carbon-Steel Castings: ASTM A 27/A 27M, Grade 60-30 (Grade 415-205).

F.

High-Strength, Low-Alloy Structural Steel: ASTM A 572/A 572M.

G.

Carbon-Steel Structural Tubing: ASTM A 500, Grade B or C.

H.

Wrought Carbon-Steel Bars: ASTM A 675/A 675M, Grade 65 (Grade 450).

I.

Deformed-Steel Wire or Bar Anchors: ASTM A 496 or ASTM A 706/A 706M.

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.6 Steel Connection Materials / 2.7 Stainless-Steel Connection Materials

6

ASTM A 307 defines the term “studs” to include stud stock and threaded rods.
J.	Carbon-Steel Bolts and Studs: ASTM A 307, Grade A or C (ASTM F 568M, Property Class 4.6) carbonsteel, hex-head bolts and studs; carbon-steel nuts (ASTM A 563/A 563M, Grade A); and flat, unhardened
steel washers (ASTM F 844).
High-strength bolts are used for friction-type connections between steel members and are not recommended
between steel and concrete because concrete creep and crushing of concrete during bolt tightening reduce
effectiveness. ASTM A 490/A 490M bolts should not be galvanized.
K.	High-Strength Bolts and Nuts: ASTM A 325/A 325M or ASTM A 490/A 490M, Type 1, heavy hex steel
structural bolts, heavy hex carbon-steel nuts, (ASTM A 563/A 563M) and hardened carbon-steel washers
(ASTM F 436/F 436M).
Structural plate and shape steel connection hardware enclosed in wall cavities is provided uncoated in non corrosive environments. Protection is required by painting or galvanizing on steel connection hardware when the corrosive environment
is high or when connections are exposed to exterior weather conditions. Retain paragraph below if shop-primed finish is
required. Indicate locations of priming, if required. MPI 79 in first option below provides some corrosion protection while
SSPC-Paint 25, without top-coating, provides minimal corrosion protection. The need for protection from corrosion will
depend on the actual conditions to which the connections will be exposed to in service.
L.

S hop-Primed Finish: Prepare surfaces of nongalvanized steel items, except those surfaces to be embedded in concrete, according to requirements in SSPC-SP 3 and shop-apply (lead- and chromate-free,
rust–inhibitive primer, complying with performance requirements in MPI 79) (SSPC-Paint 25)
according to SSPC-PA 1.

Retain paragraph and subparagraph below if galvanized finish is required. Indicate locations of galvanized
items if required. Field welding should generally not be permitted on galvanized elements, unless the galvanizing is removed or acceptable welding procedures are submitted. Hot-dip galvanized finish provides greater corrosion resistance than electrodeposited zinc coating. Electrodeposition is usually limited to threaded fasteners.
M.	Zinc-Coated Finish: For steel items in exterior walls and items indicated for galvanizing, apply zinc coating by (hot-dip process according to ASTM A 123/A 123M, after fabrication, ASTM A 153/A
153M, or ASTM F 2329 as applicable) (electrodeposition according to ASTM B 633, SC 3, Type
1and 2 and F 1941 and F 1941M).
		
1.	For steel shapes, plates, and tubing to be galvanized, limit silicon content of steel to less than
0.03% or to between 0.15 and 0.25% or limit sum of silicon content and 2.5 times phosphorous
content to 0.09%.
		
2.	Galvanizing Repair Paint: High zinc-dust-content paint with dry film containing not less than 94%
zinc dust by weight, and complying with DOD-P-21035A or SSPC-Paint 20. Comply with manufacturer’s requirements for surface preparation.

2.7 S TAINLESS-STEEL CONNECTION MATERIALS
Delete this Article if not required. Retain when resistance to staining and corrosion merits extra cost in high
moisture or corrosive areas.

ARCHITECTURAL PRECAST CONCRETE

|

561

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.7 Stainless-Steel Connection Materials / 2.8 Bearing Pads And Other Accessories

A.

Stainless-Steel Plate: ASTM A 666, Type 304, of grade suitable for application.

B.	Stainless-Steel Bolts and Studs: ASTM F 593, alloy 304 or 316, hex-head bolts and studs; stainless-steel
nuts; and flat, stainless-steel washers.
		
1.	Lubricate threaded parts of stainless-steel bolts with an anti-seize thread lubricant during assembly.
C.	Stainless-Steel Headed Studs: ASTM A 276 with the minimum mechanical properties for studs of PCI
MNL 117, Table 3.2.3.

2.8 B
 EARING PADS AND OTHER ACCESSORIES
Delete this Article if not applicable. Choice of bearing pad can usually be left to fabricator; coordinate selection
with structural engineer if required.
A.	Provide one of the following bearing pads for architectural precast concrete units (as recommended by
precast concrete fabricator for application):
		
1.	Elastomeric Pads: AASHTO M 251, plain, vulcanized, 100% polychloroprene (neoprene) elastomer,
molded to size or cut from a molded sheet, 50 to 70 Shore A durometer according to ASTM D
2240, minimum tensile strength 2250 psi (15.5 MPa) per ASTM D 412.
		
2.	Random-Oriented, Fiber-Reinforced Elastomeric Pads: Preformed, randomly oriented synthetic fibers set
in elastomer. Surface hardness of 70 to 90 Shore A durometer according to ASTM D 2240. Capable of
supporting a compressive stress of 3000 psi (20.7 MPa) with no cracking, splitting, or delaminating in
the internal portions of the pad. Test one specimen for each 200 pads used in the Project.
		
3.	Cotton-Duck-Fabric–Reinforced Elastomeric Pads: Preformed, horizontally layered cotton-duck fabric
bonded to an elastomer. Surface hardness of 80 to 100 Shore A durometer according to ASTM D
2240. Conforming to Division II, Section 18.10.2 of AASHTO LRFD Bridge Design Specifications, or
Military Specification, MIL-C-882E.
		
4.	Frictionless Pads: Tetrafluoroethylene (Teflon), glass-fiber reinforced, bonded to stainless or mild-steel
plates, or random-oriented, fiber-reinforced elastomeric pads, of type required for in-service stress.
		
5.	High-Density Plastic: Multimonomer, nonleaching, plastic strip capable of supporting loads with no
visible overall expansion.
Select material from options in paragraph below or add another material to suit Project. Coordinate with counterflashing
materials and details. It is preferable to use surface mounted reglets to avoid misalignment of reglets from panel to panel.
B.	Reglets: (PVC extrusions) (Stainless steel, Type 304) (Copper) (Reglets and flashing are specified in Division 07 Section “Sheet Metal Flashing and Trim”) felt- or fiber-filled or face opening of
slots covered.
C.	Precast Concrete Accessories: Provide clips, hangers, high-density plastic or steel shims, and other accessories required to install architectural precast concrete units.
D.	Welding Electrodes: Comply with AWS standards.

562

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.9 Grout Materials / 2.10 Clay Product Units And Accessories

6

2.9 G
 ROUT MATERIALS
Add other proprietary grout systems to suit Project. Describe locations of each grout here or on Drawings if
retaining more than one type. Indicate required strengths on Contract Drawings.
A.	Sand-Cement Grout: Portland cement, ASTM C 150, Type I, and clean, natural sand, ASTM C 144 or
ASTM C 404. Mix at ratio of 1 part cement to 21/2 to 3 parts sand, by volume, with minimum water
required for placement.
Retain first paragraph below if nonshrink grout is required or if cement-grout shrinkage could cause structural
deficiency. For critical installations, require manufacturer to provide field supervision.
B.	Nonmetallic, Nonshrink Grout: Premixed, packaged non-ferrous aggregate, noncorrosive, nonstaining
grout containing selected silica sands, portland cement, shrinkage-compensating agents, plasticizing and
water-reducing admixtures, complying with ASTM C 1107, Grade A for drypack and Grades B and C for
flowable grout and of a consistency suitable for application within a 30-minute working time.
C.	Epoxy-Resin Grout: Two-component, mineral-filled epoxy-resin: ASTM C 881/C 881M of type, grade,
and class to suit requirements.

2.10 C LAY PRODUCT UNITS AND ACCESSORIES
Retain this Article if specifying thin veneer brick–faced precast concrete panels. PCI Standard for brick units features tighter dimensional tolerances than ASTM C 1088 or ASTM C 216, Type FBX. TBX or FBX brick units may
be too dimensionally variable to fit securely within form liner templates. For economy, brick patterns should
minimize cutting of brick. Select thin brick manufacturer and product prior to bid or establish cost allowance. If
full-size brick units are required, delete this article and refer to Division 04 Section “Unit Masonry Assemblies.”
The listed characteristics for thin brick units are included in PCI “Standard for Thin Brick”.
A.	Thin Brick Units: PCI Standard, not less than 1/2 in. (13 mm), nor more than 1 in. (25 mm) thick, with an
overall tolerance of plus 0 in., minus 1/16 in. (+0 mm, -1.6 mm) for any unit dimension 8 in. (200 mm)
or less and an overall tolerance of plus 0 in., minus 3/32 in. (+0 mm, -2.4 mm) for any unit dimension
greater than 8 in. (200 mm) measured according to ASTM C 67.
		
1. Face Size: Modular, 21/4 in. (57 mm) high by 7 5/8 in. (190 mm) long.
		
2. Face Size: Norman, 21/4 in. (57 mm) high by 115/8 in. (290 mm) long.
		
3. Face Size: Closure Modular, 3 5/8 in. (90 mm) high by 7 5/8 in. (190 mm) long.
		
4. Face Size: Utility, 3 5/8 in. (90 mm) high by 115/8 in. (290 mm) long.
If approving a color range for brick, view 100 square feet (9.3 m2) of loose bricks or a completed building. Edit
to suit Project or delete if brick is specified by product name.
5.	Face Size, Color, and Texture: (Match Architect’s samples) (Match existing color, texture, and
face size of adjacent brickwork).
			
a. .
		

ARCHITECTURAL PRECAST CONCRETE

|

563

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.10 Clay Product Units And Accessories

Show details on Drawings of special conditions and shapes if required.
		
		
		
		
		
		
		
		
		
		

6. Special Shapes: Include corners, edge corners, and end edge corners.
7. Cold Water Absorption at 24 Hours: Maximum 6% when tested per ASTM C 67.
8.	Efflorescence: Tested according to ASTM C 67 and rated “not effloresced.”
9. Out of Square: Plus or minus 1/16 in. (± 1.6 mm) measured according to ASTM C 67.
10. Warpage: Consistent plane of plus 0 in., minus 1/16 in. (+0 mm, -1.6 mm).
11. Variation of Shape from Specified Angle: Plus or minus 1 degree.
12.	Tensile Bond Strength: Not less than 150 psi (1.0MPa) when tested per modified ASTM E 488.
Epoxy steel plate with welded rod on a single brick face for each test.
13.	Freezing and Thawing Resistance: No detectable deterioration (spalling, cracking, or chafing) when
tested in accordance with ASTM C 666 Method B.
14.	Modulus of Rupture: Not less than 250 psi (1.7MPa) when tested in accordance with ASTM C 67.
15.	Chemical Resistance: Provide brick that has been tested according to ASTM C 650 and rated “not affected.”

Delete subparagraph below if surface-colored brick is not used.
		

16.	Surface Coloring: Brick with surface coloring shall withstand 50 cycles of freezing and thawing per
ASTM C 67 with no observable difference in applied finish when viewed from 20 ft (6 m).

Retain first subparagraph below, deleting inapplicable descriptions if required.
		
		

17.	Back Surface Texture: scored, combed, wire roughened, ribbed, keybacked, or dovetailed.
18.	Available Products: Subject to compliance with requirements, products that may be incorporated
into the Work include, but are not limited to, the following:

Retain subparagraph above for nonproprietary or subparagraph below for semiproprietary Specification. Refer
to Division 01 Section “Materials and Equipment.”
19.	Products: Subject to compliance with requirements, products that may be incorporated into the
work include the following:
			
a.	
		

Refer to American National Standards Institute (ANSI) A 137.1 for the commonly available sizes and shapes,
physical properties, the basis for acceptance, and methods of testing.
B.
		
		
		

Glazed and Unglazed Ceramic Tile Units: ANSI A 137.1 (not less than 3/8 in. [10 mm]).
1. Body of glazed tile shall have a water absorption of less than 3% using ASTM C 373.
2. Manufacturer shall warrant materials as frost-resistant.
3. Glazed units shall conform to ASTM C 126.

C.	Architectural Terra Cotta Units: Comply with requirements of the manufacturer of the selected
Architectural Terra Cotta for the application indicated.
Retain paragraph below if mortar setting clay product unit joints before placing precast concrete mixture.
D.	Sand-Cement Mortar: Portland cement, ASTM C 150, Type I, and clean, natural sand, ASTM C 144. Mix
at ratio of 1 part cement to 4 parts sand, by volume, with minimum water required for placement.

564

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.11 Stone Materials And Accessories

6

Delete paragraph and subparagraphs below if not filling thin brick unit joints with pointing grout after precast
concrete panel production.
E.	Latex-Portland Cement Pointing Grout: ANSI A118.6 (included in ANSI A108.1) and as follows:
Select one or both types of grout from first two subparagraphs below.
		
		
		
F.

1.	Dry-grout mixture, factory prepared, of portland cement, graded aggregate, and dry, redispersible,
ethylene-vinyl-acetate additive for mixing with water; uniformly colored.
2.	Commercial portland cement grout, factory prepared, with liquid styrene-butadiene rubber or
acrylic-resin latex additive; uniformly colored.
3.	Colors: (As indicated by manufacturer’s designations) (Match Architect’s samples) (As
selected by Architect from manufacturer’s full range).
Setting Systems

Retain subparagraphs below if thin brick, ceramic tile, or full brick will be laid after casting of panel.
		
		

1.	Thin brick and Ceramic Tile Units: (Dry-Set Mortar: ANSI A118.1 [included in ANSI A108.1])
(Latex-Portland Cement Mortar: ANSI A 118.4 [included in ANSI A108.1])
2.	Full Brick Units: Install (Galvanized)(Type 304 stainless steel) dovetail slots in precast concrete:
not less than 3/16 in. (0.5 mm thick), felt- or fiber-filled or cover face opening of slots covered.
Attach brick units with wire anchors, ASTM A 82 or B 227, Grade 30HS not less than 3/16 in. (W2.8)
in diameter and hooked on one end and looped through a 7/8 in. (22 mm) wide, 12-gage (2.68 mm)
steel sheet bent over the wire with dovetail on opposite end.

2.11 STONE MATERIALS AND ACCESSORIES
Retain this Article if stone facing is required. Performance criteria, preconstruction material testing, material quality, fabrication, and finish requirements are usually specified in Division 04 Section “Exterior Stone
Cladding.” Replace first paragraph below with stone requirements, if preferred.
A.	Stone facing for architectural precast concrete is specified in Division 04 Section “Exterior Stone Cladding.”
		
1. Tolerance of length and width of +0, -1/8 in. (+0, -3 mm).
Anchors are generally supplied by stone fabricator or, in some cases, by precaster. Specify supplier. Anchors
may be toe-in, toe-out, or dowels.
B.	Anchors: Stainless steel, ASTM A 666, Type 304, of temper and diameter required to support loads without exceeding allowable design stresses.
Grommets will usually be required if filling dowel holes with rigid epoxy.
		

1.	Fit each anchor leg with 60 durometer neoprene grommet collar with a width at least twice the
diameter of the anchor and a length at least five times the diameter of the anchor.

C.	Sealant Filler: ASTM C 920, low-modulus, multicomponent, nonsag polyurethane or silicone sealant complying with requirements in Division 07 Section ”Joint Sealants” and that is nonstaining to stone substrate.

ARCHITECTURAL PRECAST CONCRETE

|

565

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.11 Stone Materials And Accessories / 2.13 Concrete Mixtures

Dowel hole filling is used to prevent water intrusion into stone and future discoloration at anchor locations.
Retain paragraph above for a flexible filler or paragraph below for a rigid filler.
D.	Epoxy Filler: ASTM C 881/C 881M, 100% solids, sand-filled non-shrinking, non-staining of type, class,
and grade to suit application.
E.	Bond Breaker: (Preformed, compressible, resilient, non-staining, non-waxing, closed-cell polyethylene
foam pad, nonabsorbent to liquid and gas, 1/8 in. [3 mm] thick) (Polyethylene sheet, ASTM D 4397, 6
to 10 mil [0.15 to 0.25 mm] thick).

2.12 I NSULATED PANEL ACCESSORIES
Retain this Article if insulated, architectural precast concrete panels are required. Specify the required thickness
for each insulation type allowed to achieve the desired aged R-value. Select insulation material from one of
three paragraphs below; if using more then one type, identify location of each on Drawings.
A.	Expanded-Polystyrene Board Insulation: ASTM C 578, Type (XI, 0.70 lb/ft3[12kg/m3]), (I, 0.90 lb/ft3
[15kg/m3])(VIII, 1.15 lb/ft3[18kg/m3])(II, 1.35 lb/ft3[22kg/m3])(IX, 1.80 lb/ft3[29 kg/m3]);
(square)(ship-lap) edges; with thickness of .
B.	Extruded-Polystyrene Board Insulation: ASTM C 578, Type (X, 1.30 lb/ft3[21kg/m3])(IV, 1.55 lb/
ft3[25 kg/m3])(VI, 1.80 lb/ft3[29 kg/m3]) (VII, 2.20 lb/ft3[35 kg/m3])(V, 3.00 lb/ft3[48 kg/m3]);
(square)(ship-lap) edges; with thickness of .
C.	Polyisocyanurate Board Insulation: Rigid, cellular polyisocyanurate thermal insulation complying with
ASTM C 591; Grade 1, Type (I, 1.8 lb/ft3[29kg/m3])(II, 2.5 lb/ft3[40kg/m3])(III, 3.0 lb/ft3[48kg/m3]);
square edged; unfaced; with thickness of .
Select wythe connectors from paragraph below.
D.	Wythe Connectors: (Glass-fiber and vinyl-ester polymer connectors), (Polypropylene pin connectors),
(Stainless-steel pin connectors), (Bent galvanized reinforcing bars) (Galvanized welded wire trusses), (Galvanized bent wire connectors) (Epoxy coated carbon fiber grid), manufactured to connect
wythes of precast concrete panels.

2.13 C ONCRETE MIXTURES
A.	Prepare design mixtures to match Architect’s sample or for each type of precast concrete required.
Revise subparagraph below if fly ash or gray silica fume are not permitted. Revise percentages to suit Project.
White silica fume is available.
		

1.	Limit use of fly ash to 20 to 40% replacement of portland cement by weight; ground granulated
blast-furnace slag to 15 to 25% of portland cement by weight; and metakaolin and silica fume to
10% of portland cement by weight.

B.	Design mixtures may be prepared by a qualified independent testing agency or by qualified precast concrete plant personnel at architectural precast concrete fabricator’s option.
C.	Limit water-soluble chloride ions to the maximum percentage by weight of cement permitted by ACI
318 (ACI 318M) or PCI MNL 117 when tested in accordance with ASTM C 1218/C 1218M.

566

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.13 Concrete Mixtures

6

Architectural precast concrete units may be manufactured with a separate “architectural” face mixture and a
“structural” backup mixture. Face and backup mixtures should have similar shrinkage and thermal coefficients
of expansion. Similar water-cementitious materials ratios and cement-aggregate ratios are recommended to
limit bowing or warping.
D.	Normalweight Concrete Face and Backup Mixtures: Proportion mixtures by either laboratory trial batch
or field test data methods according to ACI 211.1, with materials to be used on Project, to provide normalweight concrete with the following properties:
Retain subparagraph below or revise to suit Project. Higher-strength mixtures may be available; verify availability with fabricators.
		
		

1.
2.

Compressive Strength (28 Days): 5000 psi (34.5 MPa) minimum.
Release Strength: As required by design.

A maximum water-cementitious materials ratio of 0.40 to 0.45 is usual for architectural precast concrete. Lower
ratios may be possible with use of high-range water reducing admixtures. Revise ratio as required to suit Project.
		

3.

Maximum Water-Cementitious Materials Ratio: 0.45.

Water absorption indicates susceptibility to weather staining. The limit in paragraph below, corresponding to
6% by weight, is suitable for average exposures. Different parts of a single panel cannot be produced with
different absorptions. Verify that fabricator can produce units with lower water absorption because special
consolidation techniques to increase concrete density are required.
E.

Water Absorption: 6% by weight or 14% by volume, tested according to PCI MNL 117.

Lightweight backup mixtures must be compatible with normalweight face mixtures to minimize bowing or
warping. Retain paragraph below if required or as an option, if satisfactory durability and in-service performance are verified by fabricator. Coordinate with selection of normalweight face mixture option above.
F.	Lightweight Concrete Backup Mixtures: Proportion mixtures by either laboratory trial batch or field test
data methods according to ACI 211.2, with materials to be used on Project, to provide lightweight concrete with the following properties:

Retain subparagraph below or revise to suit Project. Higher-strength mixtures may be available; verify with
fabricators.
		
		

1.
2.

Compressive Strength (28 Days): 5000 psi (34.5 MPa) minimum.
Release Strength: As required by design.

Increase or decrease unit weight in subparagraph below to suit Project. Coordinate with lightweight aggregate
supplier and architectural precast concrete fabricator. Lightweight concretes with combinations of lightweight
and normalweight aggregate in mixture will usually be heavier than unit weight below.
		

3.	Unit Weight: Calculated equilibrium unit weight of 115 lb/ft3 (1842 kg/m3), where variations exceed plus
or minus 5 lb/ft3 (80 kg/m3) adjust to plus or minus 3 lb/ft3 (48 kg/m3), according to ASTM C 567.

ARCHITECTURAL PRECAST CONCRETE

|

567

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.13 Concrete Mixtures / 2.16 Stone Facings

G.	Add air-entraining admixture at manufacturer’s prescribed rate to result in concrete at point of placement having an air content complying with PCI MNL 117.
H.	When included in design mixtures, add other admixtures to concrete according to manufacturer’s written instructions.

2.14 M
 OLD FABRICATION
A.	Molds: Accurately construct molds, mortar tight, of sufficient strength to withstand pressures due to
concrete placement and vibration operations and temperature changes, and for prestressing and detensioning operations. Coat contact surfaces of molds with release agent before reinforcement is placed.
Avoid contamination of reinforcement and prestressing tendons by release agent.
Delete form liners in subparagraph below unless needed to produce exposed surface finish.
		

1.	Place form liners accurately to provide finished surface texture indicated. Provide solid backing and
supports to maintain stability of liners during concrete placement. Coat form liner with form-release
agent.

B.	Maintain molds to provide completed architectural precast concrete units of shapes, lines, and dimensions indicated, within fabrication tolerances specified.
		
1. Form joints are not permitted on faces exposed to view in the finished work.
Select one option from subparagraph below; show details on Drawings or revise description to add dimensions. Sharp edges or corners of precast concrete units are vulnerable to chipping.
		

2.

Edge and Corner Treatment: Uniformly (chamfered) (radiused).

2.15 T HIN BRICK FACINGS
Retain this Article if using thin brick facings on architectural precast concrete units.
A.	Place form liner templates accurately to provide grid for brick facings. Provide solid backing and supports
to maintain stability of liners while placing bricks and during concrete placement.
B.	Match appearance of sample panel(s).
C.	Securely place brick units face down into form liner pockets and place concrete backing mixture.
D.	After stripping units, clean faces and joints of brick facing.

2.16 S TONE VENEER FACINGS
Retain this Article if stone facing is required. Refer to Division 04 Section “Exterior Stone Cladding”.

568

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.16 Stone Facings

6

A.	Accurately position stone facings to comply with requirements and in locations indicated on Shop
Drawings. Install anchors, supports, and other attachments indicated or necessary to secure stone in
place. Maintain projection requirements of stone anchors into concrete substrate. Orient stone veining in
direction indicated on Shop Drawings. Keep reinforcement a minimum of 3/4 in. (19 mm) from the back
surface of stone. Use continuous spacers to obtain uniform joints of widths indicated and with edges
and faces aligned according to established relationships and indicated tolerances. Ensure no passage of
precast concrete matrix to stone surface.
B.	See Division 07 Section “Joint Sealants” for furnishing and installing sealant backings and sealant into
stone-to-stone joints and stone-to-concrete joints. Apply a continuous sealant bead along both sides
and top of precast concrete panels at the stone/precast concrete interface using the bond breaker as a
joint filler backer. Do not seal panel bottom edge.
Retain one of two subparagraphs below if sealing dowel holes. Use sealant if a flexible filler is required; use
epoxy if a rigid filler is required.
		
		

1. Fill anchor holes with low modulus polyurethane sealant filler and install anchors.
2.	Fill anchor holes with epoxy filler and install anchors with minimum 1/2 in. (13 mm) long, 60 durometer elastomeric sleeve at the back surface of the stone.

Retain one of two subparagraphs below. PCI recommends preventing bond between stone facing and precast
concrete to minimize bowing, cracking, and staining of stone.
		
		

3.	Install 6 to 10 mil (0.15 to 0.25 mm) thick polyethylene sheet to prevent bond between back of
stone facing and concrete substrate.
4.	Install 1/8 in. (3 mm) thick polyethylene-foam bond breaker to prevent bond between back of stone
facing and concrete substrate.

PCI recommends anchor spacing be determined prior to bidding. Retain below if precaster is to test stone
anchors for shear and tension. ASTM E 488 is preferred as ASTM C 1354 does not include the influence of the
precast concrete backup.
C.	Stone Anchor Shear and Tensile Testing: Engage accredited testing laboratory acceptable to the Architect
to evaluate and test the proposed stone anchorage system. Test for shear and tensile strength of proposed stone anchorage system in accordance with ASTM E 488 or ASTM C 1354 modified as follows:
		
1.	Prior to testing, submit for approval a description of the test assembly (including pertinent data on
materials), test apparatus, and procedures.
		
2.	Test 12 in. by 12 in. (300 mm by 300 mm) samples of stone affixed to testing apparatus through
proposed anchorages. Provide 2 sets of 6 stone samples each. One set for shear load testing and
the other set for tensile load testing.
		
3. Test stone anchors of the sizes and shapes proposed for the installation.
			
a.	Test the assembly to failure and record the test load at failure. Record the type of failure, anchor
pullout or stone breakage, and any other pertinent information, in accordance with the requirements of ASTM E 488.
Retain subparagraph below and revise anchor spacing if required as a result of preconstruction testing of stone
anchors for shear and tension specified in Division 04 Section “Exterior Stone Cladding.”

ARCHITECTURAL PRECAST CONCRETE

|

569

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.16 Stone Facings / 2.17 Fabrication

D.	Stone to Precast Concrete Anchorages: Provide anchors in numbers, types and locations required to satisfy
specified performance criteria, but not less than two anchors per stone unit of less than 2 ft2 (0.19 m2) in area
and four anchors per unit of less than 12 ft2 (1.1 m2) in area; and for units larger than 12 ft2 (1.1 m2) in area,
provide anchors spaced not more than 24 in. (600 mm) on center both horizontally and vertically. Locate anchors
a minimum of 6 in. (150 mm) from stone edge.

2.17 F ABRICATION
Coordinate with other trades for installation of cast-in items.
A.	Cast-in Anchors, Inserts, Plates, Angles, and Other Anchorage Hardware: Fabricate anchorage hardware
with sufficient anchorage and embedment to comply with design requirements. Accurately position for
attachment of loose hardware and secure in place during precasting operations. Locate anchorage hardware where it does not affect position of main reinforcement or concrete placement.
		
1.	Weld headed studs and deformed bar anchors used for anchorage according to AWS D1.1/D1.1M
and AWS C5.4, “Recommended Practices for Stud Welding.”
Coordinate paragraph below with Division 05 Section “Metal Fabrications” for furnishing and installing loose
hardware items.
B.	Furnish loose hardware items including steel plates, clip angles, seat angles, anchors, dowels, cramps,
hangers, and other hardware shapes for securing architectural precast concrete units to supporting and
adjacent construction.
C.	Cast in reglets, slots, holes, and other accessories in architectural precast concrete units as indicated on
Contract Drawings.
Delete first paragraph below if not applicable.
D.	Cast in openings larger than 10 in. (250 mm) in any dimension. Do not drill or cut openings or prestressing strand without of Architect’s approval.
E.	Reinforcement: Comply with recommendations in PCI MNL 117 for fabrication, placing, and supporting
reinforcement.
		
1.	Clean reinforcement of loose rust and mill scale, earth, and other materials that reduce or destroy
the bond with concrete. When damage to epoxy-coated reinforcing exceeds limits specified in
ASTM A 775/A 775M, repair with patching material compatible with coating material and epoxy
coat bar ends after cutting.
		
2.	Accurately position, support, and secure reinforcement against displacement during concrete- placement and
consolidation operations. Completely conceal support devices to prevent exposure on finished surfaces.
		
3.	Place reinforcing steel and prestressing tendon to maintain at least 3/4 in. (19 mm) minimum concrete cover. Increase cover requirements for reinforcing steel to 11/2 in. (38 mm) when units are
exposed to corrosive environment or severe exposure conditions. Arrange, space, and securely tie
bars and bar supports to hold reinforcement in position while placing concrete. Direct wire tie ends
away from finished, exposed concrete surfaces.
		
4.	Install welded wire reinforcement in lengths as long as practicable. Lap adjoining pieces at least one
full mesh spacing and wire tie laps, where required by design. Offset laps of adjoining widths to
prevent continuous laps in either direction.
F.	Reinforce architectural precast concrete units to resist handling, transportation and erection stresses, and
specified in-place loads, whichever governs.

570

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.17 Fabrication

6

Delete first paragraph and subparagraphs below if prestressed architectural precast concrete units are not
required. Option to prestress may be left to fabricator if objective is to aid in handling and to control cracking
of units during installation.
G.	Prestress tendons for architectural precast concrete units by pretensioning or post-tensioning methods.
Comply with PCI MNL 117.
Revise release or post-tensioning strength in subparagraph below to an actual compressive strength if required.
A concrete strength in the range of 2500 psi (17.2 MPa) to 4000 psi (27.6 MPa) at release does not appreciably
affect bond transfer length.
		

		
		
		

1.	Delay detensioning or post-tensioning of precast, prestressed architectural precast concrete units
until concrete has reached its indicated minimum design release compressive strength as established
by test cylinders cured under the same conditions as concrete member.
2.	Detension pretensioned tendons either by gradually releasing tensioning jacks or by heat-cutting
tendons, using a sequence and pattern to prevent shock or unbalanced loading.
3.	If concrete has been heat cured, detension while concrete is still warm and moist to avoid dimensional changes that may cause cracking or undesirable stresses.
4.	Protect strand ends and anchorages with bituminous, zinc-rich, or epoxy paint to avoid corrosion
and possible rust spots.

H.	Comply with requirements in PCI MNL 117 and requirements in this Section for measuring, mixing,
transporting, and placing concrete. After concrete batching, no additional water may be added.
Retain first paragraph below if a separate face mixture is required or is fabricator’s option.
I.	Place face mixture to a minimum thickness after consolidation of the greater of 1 in. (25 mm) or 1.5
times the nominal maximum aggregate size, but not less than the minimum reinforcing cover as indicated on Contract Drawings.
		
1. Use a single design mixture for those units in which more than one major face (edge) is exposed.
		
2.	Where only one face of unit is exposed, at the fabricator’s option, either of the following mixture
design/casting techniques may be used:
			
a. A single design mixture throughout the entire thickness of panel.
			
b.	Separate mixtures for face and backup concrete; using cement and aggregates for each type
as appropriate, for consecutive placement in the mold. Use cement and aggregate specified for
face mixture. Use cement and aggregate for backup mixture complying with specified criteria or
as selected by the fabricator.
J.	Place concrete in a continuous operation to prevent seams or planes of weakness from forming in
precast concrete units.
		
1. Place backup concrete to ensure bond with face-mixture concrete.
K.	Thoroughly consolidate placed concrete by internal and/or external vibration without dislocating or damaging reinforcement and built-in items, and minimize pour lines, honeycombing, or entrapped air voids
on surfaces. Use equipment and procedures complying with PCI MNL 117.
		
1.	Place self-consolidating concrete without vibration in accordance with PCI TR-6 “Interim Guidelines
for the Use of Self-Consolidating Concrete.”
L.

Comply with PCI MNL 117 procedures for hot- and cold-weather concrete placement.

M.	Identify pickup points of architectural precast concrete units and orientation in structure with permanent
markings, complying with markings indicated on Shop Drawings. Imprint or permanently mark casting
date on each architectural precast concrete unit on a surface that will not show in finished structure.
ARCHITECTURAL PRECAST CONCRETE

|

571

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.17 Fabrication / 2.19 Fabrication Tolerances

N.	Cure concrete, according to requirements in PCI MNL 117, by moisture retention without heat or by
accelerated heat curing using low-pressure live steam or radiant heat and moisture. Cure units until the
compressive strength is high enough to ensure that stripping does not have an effect on the performance or appearance of final product.
O.	Repair damaged architectural precast concrete units to meet acceptability requirements in PCI MNL 117
and Architect’s approval.

2.18 I NSULATED PANEL CASTING
Delete this Article if integrally insulated panels are not required.
A.

Cast and screed wythe supported by mold.

B.	Place insulation boards, abutting edges and ends of adjacent boards. Insert wythe connectors through
insulation, and consolidate concrete around connectors according to connector manufacturer’s written
instructions.
C.

Cast and screed top wythe to meet required finish.

2.19 F ABRICATION TOLERANCES
A.	Fabricate architectural precast concrete units of shapes, lines and dimensions indicated, so each finished
unit complies with PCI MNL 117 product tolerances as well as position tolerances for cast-in items.
Select paragraph above or first paragraph and subparagraphs below. Usually retain above unless tolerances for
Project deviate from PCI recommendations. PCI MNL 117 product tolerances, referenced above and listed below,
are standardized throughout the industry. For architectural trim units such as sills, lintels, coping, cornices,
quoins, medallions, bollards, benches, planters, and pavers, tolerances are listed in PCI MNL 135, Tolerance
Manual for Precast and Prestressed Concrete Construction.
B.	Fabricate architectural precast concrete units of shapes, lines and dimensions indicated, so each finished
unit complies with the following product tolerances.
		
1. Overall Height and Width of Units, Measured at the Face Exposed to View: As follows:
			
a. 10 ft (3 m) or under, Plus or Minus 1/8 in. (±3 mm).
			
b. 10 to 20 ft (3 to 6 m), Plus 1/8 in. (+3 mm), Minus 3/16 in. (-5 mm).
			
c. 20 to 40 ft (6 to 12 m), Plus or Minus 1/4 in. (±6 mm).
			
d. Each additional 10 ft (3 m), add Plus or Minus 1/16 in. (±1.6 mm).
		
2. Overall Height and Width of Units, Measured at the Face Not Exposed to View: As follows:
			
a. 10 ft (3 m) or under, Plus or Minus 1/4 in. (±6 mm).
			
b. 10 to 20 ft (3 to 6 m), Plus 1/4 in. (+6 mm), Minus 3/8 in. (-10 mm).
			
c. 20 to 40 ft (6 to 12 m), Plus or Minus 3/8 in. (±10 mm).
			
d. Each additional 10 ft (3 m), add Plus or Minus 1/8 in. (±3 mm).
		
3. Total Thickness or Flange Thickness: Plus 1/4 in. (+6 mm), Minus 1/8 in. (-3 mm).
		
4. Rib Width: Plus or Minus 1/8 in. (±3 mm).
		
5. Rib to Edge of Flange: Plus or Minus 1/8 in. (±3 mm).
		
6. Distance between Ribs: Plus or Minus 1/8 in. (±3 mm).
		
7.	Variation from Square or Designated Skew (Difference in Length of the Two Diagonal
Measurements): Plus or Minus 1/8 in. per 72 in. (±3 mm per 2 m) or 1/2 in. (13 mm) total, whichever is greater.

572

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.19 Fabrication Tolerances

		
		
		
		
		
		
		
		
		
		

6

8.	Length and Width of Blockouts and Openings within One Unit: Plus or Minus 1/4 in. (±6 mm).
9.	Location and Dimensions of Blockouts Hidden from View and Used for HVAC and Utility
Penetrations: Plus or Minus 3/4 in. (±19 mm).
10. Dimensions of Haunches: Plus or Minus 1/4 in. (±6 mm).
11. Haunch Bearing Surface Deviation from Specified Plane: Plus or Minus 1/8 in. (±3 mm).
12.	Difference in Relative Position of Adjacent Haunch Bearing Surfaces from Specified Relative Position:
Plus or Minus 1/4 in. (±6 mm).
13. Bowing: Plus or Minus L/360, maximum 1 in. (25 mm).
14. Local Smoothness: 1/4 in. per 10 ft (6 mm per 3 m).
15. Warping: 1/16 in. per 12 in. (1.6 mm per 300 mm) of distance from the nearest adjacent corner.
16. Tipping and Flushness of Plates: Plus or Minus 1/4 in. (±6 mm).
17. Dimensions of Architectural Features and Rustications: Plus or Minus 1/8 in. (±3 mm).

C.	Position Tolerances: For cast-in items measured from datum line location, as indicated on Shop Drawings.
		
1. Weld Plates: Plus or Minus 1 in. (±25 mm).
		
2. Inserts: Plus or Minus 1/2 in. (±13 mm).
		
3. Handling Devices: Plus or Minus 3 in. (±75 mm).
		
4.	Reinforcing Steel and Welded Wire Reinforcement: Plus or Minus 1/4 in. (±6 mm) where position has
structural implications or affects concrete cover; otherwise, Plus or Minus 1/2 in. (±13 mm).
		
5.	Reinforcing Steel Extending out of Member: Plus or Minus 1/2 in. (±13 mm) of plan dimensions.
		
6.	Tendons: Plus or Minus 1/4 in. (±6 mm), perpendicular to panel; Plus or Minus 1 in. (±25 mm),
parallel to panel.
		
7. Location of Rustication Joints: Plus or Minus 1/8 in. (±3 mm).
		
8. Location of Opening within Panel: Plus or Minus 1/4 in. (±6 mm).
		
9. Location of Flashing Reglets: Plus or Minus 1/4 in. (±6 mm).
		
10. Location of Flashing Reglets at Edge of Panel: Plus or Minus 1/8 in. (±3 mm).
		
11. Reglets for Glazing Gaskets: Plus or Minus 1/8 in. (±3 mm).
		
12. Electrical Outlets, Hose Bibs: Plus or Minus 1/2 in. (±13 mm).
		
13. Location of Bearing Surface from End of Member: Plus or Minus 1/4 in. (±6 mm).
		
14.	Allowable Rotation of Plate, Channel Inserts, Electrical Boxes: 2-degree rotation or 1/4 in. (6 mm)
maximum measured at perimeter of insert.
		
15.	Position of Sleeve: Plus or Minus 1/2 in. (±13 mm).
		
16. Location of Window Washer Track or Buttons: Plus or Minus 1/8 in. (±3 mm).
Delete paragraph below if brick faced architectural units are not used. The number of bricks allowed these
misalignments should be limited to 2% of the bricks on the unit.
D. Brick-Faced Architectural Precast Concrete Units.
		
1. Alignment of mortar joints:
			
a. Jog in Alignment: 1/8 in. (3 mm).
			
b. Alignment with Panel Centerline: Plus or Minus 1/8 in. (±3 mm).
		
2. Variation in Width of Exposed Mortar Joints: Plus or Minus 1/8 in. (±3 mm).
		
3.	Tipping of Individual Bricks from the Panel Plane of Exposed Brick Surface: Plus 0 in. (+0 mm);
Minus 1/4 in. (-6 mm) ≤ depth of form liner joint.
		
4.	Exposed Brick Surface Parallel to Primary Control Surface of Panel: Plus 1/4 in. (+6 mm); Minus 1/8 in.
(-3 mm).
		
5.	Individual Brick Step in Face from Panel Plane of Exposed Brick Surface: Plus 0 in. (+0 mm); Minus
1
/4 in. (-6 mm) ≤ depth of form liner joint.

ARCHITECTURAL PRECAST CONCRETE

|

573

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.19 Fabrication Tolerances / 2.20 Finishes / 2.21 Source Quality Control

Delete paragraph and subparagraphs below if stone veneer–faced architectural precast concrete units are not used.
E.

Stone Veneer–Faced Architectural Precast Concrete Units.

Tolerances below are generally appropriate for smooth-finished stone. Retain, delete, or revise to suit Project.
		
		

1.	Variation in Cross-Sectional Dimensions: For thickness of walls from dimensions indicated: Plus or
Minus 1/4 in. (±6 mm).
2.	Variation in Joint Width: 1/8 in. in 36 in. (3 mm in 900 mm) or a quarter of nominal joint width,
whichever is less.

Revise or delete below for natural-cleft, thermal, and similar finishes.
		

3.	Variation in Plane between Adjacent Stone Units (Lipping): 1/16 in (1.6 mm) difference between
planes of adjacent units.

2.20 F INISHES
A.	Exposed panel faces shall be free of joint marks, grain, and other obvious defects. Corners, including
false joints shall be uniform, straight, and sharp. Finish exposed-face surfaces of architectural precast
concrete units to match approved (design reference sample) (sample panels) (mockups) and as
follows:
This Article presumes Architect has preapproved one or more design reference samples. Include complete
description of design reference sample here. If preapproving fabricators, coordinate with “Fabricators” Article.
Revise if multiple samples are approved.
		

1.	Design Reference Sample: 

Delete subparagraph below if not required. PCI published numbered, color photographs of 428 precast concrete finishes. See PCI’s website at www.pci.org for more information. If retaining, revise and add reference
number. Add reference number combinations if more than one finish is required.
		

2.	PCI’s Architectural Precast Concrete –Color and Texture Selection Guide, of plate numbers indicated.

Select type of finish from subparagraphs below if needed. If more than one finish is required, add locations to
finish descriptions or indicate on Drawings. Add more detailed descriptions of finishes outlined below if greater
definition is required, such as (light), (medium), or (deep). Remove matrix to a maximum depth of one-third
the average diameter of coarse aggregate but not more than one-half the diameter of smallest-sized coarse
aggregate. See PCI MNL 117 for more information on finishes. An as-cast finish generally results in a mottled
surface or non-uniform finish.
		
		
		
		

574

|

3.	As-Cast Surface Finish: Provide surfaces free of excessive air voids, sand streaks, and honeycombs.
4.	Textured-Surface Finish: Impart texture by form liners with surfaces free of excessive air voids, sand
streaks, and honeycombs, with uniform color and texture.
5.	Bushhammer Finish: Use power or hand tools to remove matrix and fracture coarse aggregates.
6.	Exposed Aggregate Finish: Use chemical retarding agents applied to molds, and washing and brushing procedures, to expose aggregate and surrounding matrix surfaces after form removal.

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

2.20 Finishes / 2.21 Source Quality Control

		
		

		
		
		
		
		

6

7.	Abrasive-Blast Finish: Use abrasive grit, equipment, application techniques, and cleaning procedures
to expose aggregate and surrounding matrix surfaces.
8.	Acid-Etched Finish: Use acid and hot-water solution, equipment, application techniques, and cleaning procedures to expose aggregate and surrounding matrix surfaces. Protect hardware, connections, and insulation from acid attack.
9.	Honed Finish: Use continuous mechanical abrasion with fine grit, followed by filling and rubbing
procedures.
10.	Polished Finish: Use continuous mechanical abrasion with fine grit, followed by filling and rubbing
procedures.
11.	Sand-Embedment Finish: Use selected stones placed in a sand bed in bottom of mold, with sand
removed after curing.
12.	Thin Brick Facings: Refer to “Thin Brick Facings” Article.
13.	Stone Veneer Facings: Refer to “Stone Veneer Facings” Article.

B.	Finish exposed (top) (bottom) (back) surfaces of architectural precast concrete units to match facesurface finish.
Revise finish in paragraph below to light-broom, stippled, or float finish, if necessary. Upgrade to steel-trowel
finish if surface is in contact with materials requiring a smooth finish.
C.

Finish unexposed surfaces of architectural precast concrete units with as-cast finish.

Retain paragraph above or below if applicable. Revise below to float finish or light-broom finish if steel-trowel
finish is unnecessary.
D.	Finish unexposed surfaces (top) (back) of architectural precast concrete units by steel-trowel finish.

2.21 S OURCE QUALITY CONTROL
Always retain paragraph below because it establishes a minimum standard of plant testing and inspecting. PCI
MNL 117 mandates source testing requirements and a plant “Quality Systems Manual.” PCI certification also
ensures periodic auditing of plants for compliance with requirements in PCI MNL 117.
A.	Quality-Control Testing: Test and inspect precast concrete according to PCI MNL 117 requirements. If
using self-consolidating concrete also test and inspect according to PCI TR-6 “Interim Guidelines for the
Use of Self-Consolidating Concrete.”
Delete first paragraph and subparagraph below if not required. PCI certification would normally be acceptable to
authorities having jurisdiction without further monitoring of plant quality control and testing program by Owner.
B.	In addition to PCI Certification, Owner will employ an accredited independent testing agency to evaluate
architectural precast concrete fabricator’s quality-control and testing methods.
		
1.	Allow Owner’s testing agency access to material storage areas, concrete production equipment, and
concrete placement and curing facilities. Cooperate with Owner’s testing agency and provide samples
of materials and concrete mixtures as may be requested for additional testing and evaluation.
C.	Strength of precast concrete units will be considered deficient if units fail to comply with ACI 318 (ACI
318M) concrete strength requirements.

ARCHITECTURAL PRECAST CONCRETE

|

575

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
2.21 Source Quality Control / 3.2 Examination

Review testing and acceptance criteria with structural engineer. In first paragraph and subparagraphs below,
add criteria for load tests if required.
D.	Testing: If there is evidence that strength of precast concrete units may be deficient or may not comply with ACI
318 (ACI 318M) requirements, fabricator will employ an independent testing agency to obtain, prepare, and test
cores drilled from hardened concrete to determine compressive strength according to ASTM C 42/C 42M.
		
1.	A minimum of three representative cores will be taken from units of suspect strength, from
locations directed by Architect.
		
2. Cores will be tested in an air-dry condition.
		
3.	Strength of concrete for each series of three cores will be considered satisfactory if the average
compressive strength is equal to at least 85% of the 28-day design compressive strength and no
single core is less than 75% of the 28-day design compressive strength.
		
4.	Test results will be reported in writing on the same day that tests are performed, with copies to
Architect, Contractor, and precast concrete fabricator. Test reports will include the following:
			
a. Project identification name and number.
			
b. Date when tests were performed.
			
c. Name of precast concrete fabricator.
			
d. Name of concrete testing agency.
			
e.	Identification letter, name, and type of precast concrete unit(s) represented by core tests; design
compressive strength; type of break; compressive strength at breaks, corrected for length-diameter
ratio; and direction of applied load to core in relation to horizontal plane of concrete as placed.
E.	Patching: If core test results are satisfactory and precast concrete units comply with requirements, clean
and dampen core holes and solidly fill with precast concrete mixture that has no coarse aggregate, and
finish to match adjacent precast concrete surfaces.
F.	Defective Work: Architectural precast concrete units that do not comply with acceptability requirements in
PCI MNL 117, including concrete strength, manufacturing tolerances, and color and texture range are unacceptable. Chipped, spalled, or cracked units may be repaired, if repaired units match the visual mock-up.
The Architect reserves the right to reject any unit if it does not match the accepted sample panel or visual
mock-up. Replace unacceptable units with precast concrete units that comply with requirements.

PART 3 – EXECUTION
3.1 P
 REPARATION
A.	Deliver anchorage devices for precast concrete units that are embedded in or attached to the building
structural frame or foundation before start of such work. Provide locations, setting diagrams, and templates for the proper installation of each anchorage device.

3.2 E XAMINATION
A.	Examine supporting structural frame or foundation and conditions for compliance with requirements for
installation tolerances, true and level bearing surfaces, and other conditions affecting precast concrete
performance.
B.	Proceed with precast concrete installation only after unsatisfactory conditions have been corrected.
C.	Do not install precast concrete units until supporting cast-in-place concrete building structural framing
has attained minimum allowable design compressive strength or supporting steel or other structure is
structurally ready to receive loads from precast concrete units.

576

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
3.3 Erection

6

3.3 E RECTION
A.	Install loose clips, hangers, bearing pads, and other accessories required for connecting architectural
precast concrete units to supporting members and backup materials.
Retain one of two paragraphs below
B.	Structural steel fabricator to supply and install miscellaneous steel preweld connection hardware in the shop.
C.	Precaster or erector to supply and install miscellaneous steel preweld connection hardware in the field.
D.	Erect architectural precast concrete level, plumb, and square within the specified allowable erection
tolerances. Provide temporary supports and bracing as required to maintain position, stability, and alignment of units until permanent connections are completed.
		
1.	Install steel or plastic spacing shims as precast concrete units are being erected. Tack weld steel
shims to each other to prevent shims from separating.
		
2. Maintain horizontal and vertical joint alignment and uniform joint width as erection progresses.
		
3.	Remove projecting lifting devices and use sand-cement grout to fill voids within recessed lifting
devices flush with surface of adjacent precast concrete surfaces when recess is exposed.
		
4.	Unless otherwise indicated, provide for uniform joint widths of 3/4 in. (19 mm).
E.	Connect architectural precast concrete units in position by bolting, welding, grouting, or as otherwise
indicated on Shop (Erection) Drawings. Remove temporary shims, wedges, and spacers as soon as practical after connecting and/or grouting are completed.
		
1.	Disruption of roof flashing continuity by connections is not permitted; concealment within roof
insulation is acceptable.
F.	Welding: Comply with applicable AWS D1.1/D1.1M and AWS D1.4 requirements for welding, welding
electrodes, appearance of welds, quality of welds, and methods used in correcting welding work.
		
1.	Protect architectural precast concrete units and bearing pads from damage during field welding or
cutting operations and provide noncombustible shields as required.
		
2.	Welds not specified shall be continuous fillet welds, using not less than the minimum fillet as
specified by AWS.
		
3.	Clean weld-affected metal surfaces with chipping hammer followed by brushing and then reprime
damaged painted surfaces in accordance with paint manufacturer’s recommendations.
Retain last subparagraph above or first subparagraph below.
		

		

4.	Clean weld-affected metal surfaces with chipping hammer followed by brushing and then apply a
minimum 0.004-in.-thick (0.1 mm) coat of galvanized repair paint to galvanized surfaces in conformance with ASTM A 780.
5.	Visually inspect all welds critical to precast concrete connections. Visually check all welds for
completion and remove, reweld or repair all defective welds, if services of AWS-certified welding
inspector are not furnished by Owner.

G.	At bolted connections, use lock washers, tack welding, or other approved means to prevent loosening
of nuts after final adjustment.
		
1.	Where slotted connections are used, check bolt position and tightness. For sliding connections,
properly secure bolt but allow bolt to move within connection slot. For friction connections, apply
specified bolt torque and check 25% of bolts at random by calibrated torque wrench.

ARCHITECTURAL PRECAST CONCRETE

|

577

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
3.3 Erection / 3.5 Field Quality Control

In paragraph below revise locations and extent of grouting if required.
H.	Grouting or Dry-Packing Connections and Joints: Indicate joints to be grouted and any critical grouting
sequences on Shop (Erection) Drawings. Grout connections where required or indicated on Shop (Erection)
Drawings. Retain flowable grout in place until strong enough to support itself. Alternatively pack spaces
with stiff dry pack grout material, tamping until voids are completely filled. Place grout and finish smooth,
level, and plumb with adjacent concrete surfaces. Promptly remove grout material from exposed surfaces
before it affects finishes or hardens. Keep grouted joints damp for not less than 24 hours after initial set.

3.4 E RECTION TOLERANCES
A.	Erect architectural precast concrete units level, plumb, square, true, and in alignment without exceeding
the noncumulative erection tolerances of PCI MNL 117, Appendix I.
Select paragraph above or paragraph and subparagraphs below. Usually retain above unless tolerances for
Project deviate from PCI recommendations. PCI MNL 117 erection tolerances are referenced above and are
listed below. If tighter tolerances are required for Project, coordinate with fabrication tolerances for precast
concrete as well as erection tolerances for supporting construction.
B.	Erect architectural precast concrete units level, plumb, square, and true, without exceeding the following
noncumulative erection tolerances.
		
1. Plan Location from Building Grid Datum: Plus or Minus 1/2 in. (±13 mm).
		
2. Plan Location from Centerline of Steel Support: Plus or Minus 1/2 in. (±13 mm).
		
3. Top Elevation from Nominal Top Elevation:
			
a. Exposed Individual Panel: Plus or Minus 1/4 in. (±6 mm).
			
b. Non-Exposed Individual Panel: Plus or Minus 1/2 in. (±13 mm).
		
4. Support Elevation from Nominal Support Elevation:
			
a. Maximum Low: 1/2 in. (13 mm).
			
b. Maximum High: 1/4 in. (6 mm).
		
5. Maximum Plumb Variation over the Lesser of Height of Structure or 100 ft (30 m): 1 in. (25 mm).
		
6. Plumb in Any 10 ft (3 m) of Element Height: 1/4 in. (6 mm).
		
7. Maximum Jog in Alignment of Matching Edges:
			
a. Exposed Panel Relative to Adjacent Panel: 1/4 in. (6 mm).
			
b. Non-Exposed Panel Relative to Adjacent Panel: 1/2 in. (13 mm).
		
8. Joint Width (Governs over Joint Taper): Plus or Minus 1/4 in. (±6mm).
		
9. Maximum Joint Taper: 3/8 in. (10 mm).
		
10. Joint Taper over 10 ft (3 m): 1/4 in. (6 mm).
		
11. Maximum Jog in Alignment of Matching Faces: 1/4 in. (6 mm).
		
12.	Differential Bowing or Camber, as Erected, between Adjacent Members of Same Design: 1/4 in. (6 mm).
		
13. Opening Height between Spandrels: Plus or Minus 1/4 in. (± 6 mm).

3.5 F IELD QUALITY CONTROL
Retain first option in paragraph below if Owner engages a special inspector. If authorities having jurisdiction
permit Contractor to engage a special inspector, retain second opotion and retain option for submitting special
inspection reports in Part 1 “Submittals” Article.

578

|

ARCHITECTURAL PRECAST CONCRETE

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE

3.5 Field Quality Control / 3.6 Repairs

6

A.	Special Inspections: [Owner will engage][Contractor will engage] a qualified special inspector to
perform the following special inspections and prepare reports:
		
1.	Erection of loadbearing precast concrete members.
		
2.	
Retain first paragraph below if field testing and inspecting are required, with or without paragraph above, to
identify who shall perform tests and inspections. If retaining second option, retain requirement for field qualitycontrol test reports in Part 1 “Submittals” Article.
B.	Testing: Owner will engage accredited independent testing and inspecting agency to perform field tests
and inspections and prepare reports.
		
1.	Field welds will be subject to visual inspections and nondestructive testing in accordance with ASTM
E165 or ASTM E 709.
		
2.	Testing agency will report test results promptly and in writing to Contractor and Architect.
C.	Repair or remove and replace work where tests and inspections indicate that it does not comply with
specified requirements.
D.	Additional testing and inspecting, at Erector’s expense, will be performed to determine compliance of
corrected work with specified requirements.

3.6 R
 EPAIRS
Production chips, cracks, and spalls should have been corrected at manufacturer’s plant. Blemishes occurring
after delivery are normally repaired before final joint sealing and cleaning as weather permits.
A.

Repairs will be permitted provided structural adequacy of units and appearance are not impaired.

B.

Repair damaged units to meet acceptability requirements of PCI MNL 117.

The precast concrete fabricator should develop appropriate repair mixtures and techniques during the production sample approval process.
C.	Mix patching materials and repair units so cured patches blend with color, texture, and uniformity of adjacent exposed surfaces and show no apparent line of demarcation between original and repaired work,
when viewed in typical daylight illumination from a distance of 20 ft (6 m).
D.	Prepare and repair damaged galvanized coatings with galvanizing repair paint according to
ASTM A 780.
Retain paragraph above if using galvanized anchors, connections, and other items; retain first paragraph below
if items are prime painted.
E.

Wire brush, clean, and paint damaged prime-painted components with same type of shop primer.

F.	Remove and replace damaged architectural precast concrete units when repairs do not comply with
specified requirements.

ARCHITECTURAL PRECAST CONCRETE

|

579

6

GUIDE SPECIFICATION FOR ARCHITECTURAL PRECAST CONCRETE
3.7 Cleaning

3.7 C LEANING
Specify whether erector or precaster does cleaning under the responsibility of General Contractor.
A.

Clean all surfaces of precast concrete to be exposed to view, as necessary, prior to shipping.

B.	Clean mortar, plaster, fireproofing, weld slag, and any other deleterious material from concrete
surfaces and adjacent materials immediately.
C.	Clean exposed surfaces of precast concrete units after erection and completion of joint treatment to
remove weld marks, dirt, stains and other markings.
		
1.	Perform cleaning procedures, if necessary, according to precast concrete fabricator’s recommendations. Protect adjacent work from staining or damage due to cleaning operations.
		
2.	Do not use cleaning materials or processes that could change the appearance of exposed concrete
finishes or damage adjacent materials.
END OF SECTION 034500

580

|

ARCHITECTURAL PRECAST CONCRETE

INDEX
BY SUBJ ECTS
For ease of reference and usage of this Manual an Index of the most common terms and key words follows:

A

Absorption . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 245, 254-256, 567
Acceptability . .  . 99, 107, 156, 235-236, 240-241, 266
Access . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267, 290-291
Acid-Etching. .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Acoustical Properties . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 487-497
Composite Wall Considerations . .  .  .  .  .  .  .  .  . 493-496
Definitions. .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 487-488
Leaks and Flanking . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 496-497
Noise Criteria . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 491-493
Sound Absorption . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 489, 491
Sound Transmission . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 488-489
Windows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 493-496
Admixtures . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 559-560
Advantages . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Benefits
Aggregates . .  .  .  .  .  . 100-101, 148-150, 182, 559-560
Continuously Graded . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 111
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Gap Graded . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 111
Polishing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Size . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 151
Aggregate Transparency . .  .  .  .  .  .  .  .  .  .  .  .  .  . 154-155
Air-Barrier . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 423, 444-447
Air-Entrainment . .  .  .  .  .  .  .  .  .  .  .  .  .  . 110, 254, 256, 568
Air Holes. .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Bug/Blow Holes
Air Infiltration (Leakage) . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 484
Air Pollutants . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 243
Air Traffic Control Towers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 28
Air Voids . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Bug/Blow Holes
Albedo . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393-394, 475-476
Alignment Connection . .  .  .  .  .  .  .  .  .  . see Connections
Anchors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 204-205
Anti-Graffiti Coatings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 258
Appearance Uniformity . .  . 59-60, 99-100, 109, 153,
157, 169, 174, 182, 184, 211
Applications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 2-35
Cladding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 2-23, 45-53, 63-69
Loadbearing Units . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 4-6, 8, 83-91
Miscellaneous . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 24-35
Approval . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .
Samples . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 102-109
Production Units . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 104-108
Shop Drawings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 299-300
Architectural Trim Units . .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Arrises . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 125, 128-129
Art and Sculpture . .  .  .  .  .  .  .  .  .  .  .  .  .  . 24-26, 30-31, 34

B

Backup Mixture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109-110
Balustrades . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 32

Bearing (Direct and Eccentric)
Connections. .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Connections
Bearing Pads . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 322, 324, 562
Beam Rotation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 283-284
Bell Towers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 28-30
Benefits . .  .  .  .  .  .  .  .  .  .  .  .  . 6, 8, 13, 15, 21, 35-38, 211
Blast Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 497-525
Analyses Standards . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 499-500
Basics . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 498-499
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 314-315, 517-520
Design Concepts . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 506, 508-509
Effects . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 501-504
Façade Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 509
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 497-498
Glazing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 520-523
Initial Costs . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 523-524
Loading . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 500-501
Modelling . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 501-502
Panel Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 509-511, 513
References . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 524-525
Reinforcement Requirements . .  .  .  .  .  .  .  .  .  .  . 508-509
Standoff Distance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 504-507
Bowing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 278-279, 297, 302
Sandwich Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 458-459
Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 348-350
Bracing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 293, 299, 323
Brick Facing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Bug/Blow Holes . .  .  .  .  .  .  .  .  .  . 139, 153, 155, 169, 240
Building Shape . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 498
Building Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 356-358
Bullnoses . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 125-127
Bushhammering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes

C

Carbonation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 307-309
Cast Stone . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Cantilevered Supports . .  .  .  .  .  .  .  .  .  .  .  .  . 70, 283-284
Cast-in-Place Concrete Frame Tolerances . . 357-358
Cellular Communication Towers . .  .  .  .  .  .  .  .  .  . 28-29
Cement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100, 254, 558-559
Color . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 144-145
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59
Ceramic Tile . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Certification
Field . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 38-39, 553
Personnel . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 38-39
Plant . . . . . . . . . . . . . . . . . . . . . . . . 36, 38, 270, 553
Chamfers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 58, 207-208
Cladding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 62-73
Bowing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278-279
Column Covers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 280-281

ARCHITECTURAL PRECAST CONCRETE

|

581

INDEX BY SUBJECTS
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 280, 316
Deflection of Support . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 277-278
Design Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 276-281
Spandrels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 282-285
Clay Product-Faced Precast Concrete. .  .  . see Finishes
Cleaning . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 259-261, 297-298, 580
Clearance . .  .  .  .  .  .  .  .  .  .  .  . 279-280, 320-321, 362-363
Closed Shapes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 111-112, 379-380
Coarse Aggregates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100, 148-150
Aggregate Transparency . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 154-155
Coefficients of Thermal Expansion . .  .  .  .  .  .  .  .  .  . 214
Color
Aggregate, Effects on . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 148-150
Control . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100-101
Finishes, Effect on . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 149
Pigments . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 145-148
Texture, Effect on . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 144
Uniformity . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 145, 148, 150, 153
Column Covers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 68, 138, 280-281
Clearance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 513
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 332, 340-341
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Insulation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 281
Columns . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 27-28, 30-31
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 296
Concrete . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .
Air-Entrained . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 110, 254, 256, 568
Assessment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109-111
Backup Mixtures . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109-110
Carbonation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 307-309
Cover . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 184, 306-307, 310, 570
Design Strength . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 110, 567
Durability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 35, 110, 255-256
Face Mixtures . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109-110
Fire Endurance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 535
Lightweight Concrete . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 567
Mixtures . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61, 566-568
Permeability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 307
Strength . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 110, 567
Weathering. . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 255-256
Condensation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 408, 417-447
Air Barriers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 444-447
Climate, Effect on . .  .  .  .  . 418-419, 428-429, 446-447
Dew Point Analysis . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 441-444
Moisture Sources . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 418, 420
On Surface . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 420-423
Vapor Retarders . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 423-444
Within Wall . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 423, 442-444
Windows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 447
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267, 312-346
Alignment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 332, 339-340
Anchorage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 321-322
Bearing (Direct and Eccentric) . .  .  .  .  .  .  .  .  .  .  . 331-335
Blast Considerations . .  .  .  .  .  .  .  .  .  .  . 314-315, 517-520
Bolted . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 329
Chemical Anchors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330
Clearance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 326-327
Column Covers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 332, 340-341

582

|

ARCHITECTURAL PRECAST CONCRETE

Connection-Structure Interaction . .  .  .  .  .  .  .  . 316-320
Corrosion Protection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 326-327
Design Considerations . .  .  .  .  .  .  .  .  . 313-315, 321-322
Design Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 313
Details . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 331-346
Erection Considerations . .  .  .  .  .  .  .  .  .  .  . 297, 322-324
Expansion Anchors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 329-330
Fastening Methods . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 327-330
Fireproofing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 327-328, 535-536
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 312-313
Grouted . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330
Handling and Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 322-324
Hardware and Materials . .  .  .  .  . 61, 325-326, 560-562
Locations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 315-319
Loads . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 274
Manufacturing Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  . 325
Post-Tensioned . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330
Preweld . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Seismic Considerations . .  .  .  .  .  .  .  .  .  .  . 314, 316-320
Shear Plates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 322, 332, 338-339
Slab to Wall . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 333, 345-346
Soffit Hanger . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 332, 342
Special Conditions and Solutions . .  .  .  . 332, 342-343
Standardization . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 324, 329
Supply . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330-331, 545-546
Tieback (Lateral) . .  .  .  .  .  .  .  .  .  .  .  .  . 321, 332, 336-338
Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 320-321, 357
Wall to Foundation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 333, 344
Wall to Wall . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 333, 339-340, 346
Welded . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 328-329
Wind Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 314
Construction Manager Responsibility . .  .  .  . 267-269
Continuously Graded Aggregate . .  .  .  .  .  .  .  .  .  .  .  . 111
Contract Documents . .  .  .  .  .  .  .  .  .  . 298-300, 313, 545
Estimates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 45
Coordination . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 545-546
Copings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 540-541
Corners . .  .  .  .  . 131, 133-138, 183, 190-191, 202-203
Sandwich Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 459
Volumes Changes, Effect on . .  .  .  .  .  .  .  .  .  .  .  . 276-277
Cornices . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 27-28, 130-132
Corrosion Protection . .  .  .  .  .  .  .  .  .  .  . 310-312, 326-327
Corrosion Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 305-312
Cost
Aggregates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Cement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 58, 290
Face Mixture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59
Factors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 45-50, 53, 58, 60
Finishes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Handling . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 140-141
Hardware . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Initial . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 36
Labor . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59, 61
Life Cycle . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 36-37
Loadbearing Units . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 74
Material . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Molds . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 54-55, 121

Pigment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Sculpturing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 121
Total Wall Analysis . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 62
Transportation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 143
Cover . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 184, 306-307, 310, 570
Cracks . .  .  .  .  .  .  . 241, 300, 302, 309-310, 312, 328-329
Crazing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 155, 241
Creep . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 276
Curtain Walls . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Cladding

D

Damage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Repair
Definitions . .  .  .  .  . 39-43, 312-313, 390-391, 459-460,
487-488, 548
Deflection of Supporting Structure . .  . 277-278, 321
Degrees of Exposure . .  . 152, 159, 169, 174-175, 182
Demarcation Features . .  .  .  .  . 113-119, 177, 179, 210
Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .
Checklist . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 273-274
Cladding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 276-287
Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 1, 81-83 276-285
Criteria . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 272-273
Economy . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 50, 271, 314
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 271
Loads . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 274-275
Loadbearing Units . .  .  .  .  .  .  .  .  .  .  .  . 281-282, 285-286
Objectives . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 271-272
Options . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61-62
Reference Sample . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 263-267, 269-270
Seismic . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 272
Structural . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 271-276
Detailing Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 207
Dew Point Analysis . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 441-444
Dimensioning . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 287, 289
Dissimilar Metals . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 325
Domes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 2-3, 8, 14-15
Dowels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330
Draft . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 56, 108-109, 112-113
Drift . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 314, 316-319
Drips . .  .  .  .  .  .  .  .  .  .  .  . 244, 246-247, 250, 380-382, 390
Ductility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 94, 313
Durability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Concrete

E

Economy, Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 50, 271, 314
Edges . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 131, 133, 207
Efflorescence . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 148, 250-255
Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259-260
Energy Conservation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 390-417
Albedo . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393-394
Building Orientation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 392
Building Shape . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393
Daylighting . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393
Definitions . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 390-391
Glazing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393

Energy Efficiency . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 37
Heat Capacity . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 410-412
Insulation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 447-450, 455
Sandwich Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 450-459
Shading . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 394-403
Thermal Bridges . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 407-410, 422
Thermal Mass . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 410, 412-417
Thermal Resistance (R-value) . .  .  .  .  .  .  .  . 402, 404-410
Energy Star . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 470-472
Envelope Mold . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 56-57
Environmental Impact . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 37
Epoxy-Coated Reinforcement . .  .  .  .  .  .  .  .  .  . 310-312
Erection
Access . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267, 290-291
Bracing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 293, 299, 323
Columns . . . . . . . . . . . . . . . . . . . . . . . . 296, 533-535
Column Covers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297, 533-535
Connections: see also Connections . .  .  .  .  .  . 290-295
Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 290-298, 322-324
Drawings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 293, 298, 330
Mullions . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Protection During . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 292-293
Soffit . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Spandrels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Specifications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 577-578
Structural Limitations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 294-295
Survey . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 270
Tolerance . .  .  .  .  .  .  .  .  .  .  .  .  .  . 235, 350, 356-363, 578
Wall Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 295-296
Expansion Anchors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 329-330
Expansion Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 366-368
Exposed Aggregate Finish . .  .  . see Finishes, Retarded
Exposure Differences . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 138-139
Eyebrows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 130-131

F

Façade Patterns . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 63
Face Mixture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109-110
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59
False Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 53, 58-59, 113-118
Fine Aggregates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100, 148, 254
Finishes
Acceptability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 240-241
Acid-Etching . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 174-176
Architectural Trim Units (Cast Stone) . .  .  .  .  .  . 231-237
		 Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 234
Brick . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Clay Product-Faced
Bushhammering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 182-184
Ceramic Tile . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 197-199, 203-204
	Clay Product-Faced
Precast Concrete . .  .  .  .  . 187-206, 563-564, 567, 573
		 Benefits . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 188
		 Bond . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 200-201
		 Clay Product Selection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 190
		 Corner Details . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 202-203
General Considerations . .  .  .  .  .  .  . 188, 190-191, 568

ARCHITECTURAL PRECAST CONCRETE

|

583

INDEX BY SUBJECTS

		 Patterns . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 190
		 PCI Thin Brick Standard . .  .  .  .  .  .  .  .  . 189, 563-564
		 Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 203, 573
Color . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 144-150
		 Finishing, Effect on . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 149
Combination of . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 177-181
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-61
Demarcation Features . .  .  .  .  . 53, 58-59, 113-118, 125
Exposed Aggregate Finishes . .  .  .  .  . see specific Finish
Exposure . .  .  .  .  .  .  .  .  .  . 152, 159, 169, 174-175, 182
Form Liners . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 160-167
Fractured Fin . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 185
Full Brick . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 204-206
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 153
Hammered Rib . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 185
Honed . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 206-210
Interior . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 239-240
Matching Precast and Cast-In-Place Concrete . . 238-239
	Multiple Mixtures and
Textures Within a Single Unit . .  .  .  .  .  .  .  .  .  .  . 177-181
Painted . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 229-232
Polishing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206-210
Retarded . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 156-159
Returns . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 138-139
Sand or Abrasive Blasting . .  .  .  .  .  .  .  .  .  .  .  .  .  . 169-174
Sand Embedment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 186-187
Smooth As-Cast . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 101, 153
Specifications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 574-575
Stone Veneer-Faced Precast Concrete . .  .  .  .  . 211-229
		 Accents or Feature Strips. .  .  .  .  .  .  .  .  . 218, 227-229
		 Anchorage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 214-217, 565, 570
		 Applications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 218-229
		 Bondbreaker . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 214-215, 566
		 Jointing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 217
		 Permeability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 213
		 Properties . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 212, 569
		 Safety Factors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 217
		 Sizes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 213
		 Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 213, 574
Terra Cotta . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 199-201
Texture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 144, 150-152
Tooling . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Bushhammering
Weathering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 248-250
Fire Barriers (Safing) . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 532-533
Fire Endurance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 527-532
Aggregate, Effect of . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 527-529
Columns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533
Column Covers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 533-535
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 535-536
Reinforcing Steel . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 535
Ribbed Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 529-530
Sandwich Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 530
Window Walls . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 530-532
Fire Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 37, 525-536
Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 205-206, 234, 539-540
Floor-to-Floor Height . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 74-75, 83

584

|

ARCHITECTURAL PRECAST CONCRETE

Formliners
Types . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 160, 167
Selection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 160
Size . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 161
Texture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 161
Formwork, Precast Concrete used as . .  .  .  .  .  .  . 94-97
Fractured Fin . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Frame Shortening . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 279-280
Freeze-Thaw Tests . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 256
Full Brick
Anchors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 204-205
Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 205-206
Shelf Angles . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 205
Weep Holes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 206

G

Galvanized Reinforcement . .  .  .  .  .  . 304-305, 310-311
Gap Graded Aggregates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 111
General Contractor/CM Responsibility . .  .  . 267-269
Glass Staining or Etching . .  .  .  .  .  .  .  .  .  .  .  .  .  . 387-390
Glazing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Windows
Graffiti Repellents . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 258
Green Globes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 472-473
Ground Floors, Precast Concrete Units for . .  .  .  .  . 46
Grouting . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 330

H

Hammered Rib . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Handling of Precast Concrete Units . .  . 290-292, 322,
324-325
Heat Capacity . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 410-412
Heat Transmittance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see U Values
Honed . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Hybrid Moment Resistant Frame . .  .  .  .  .  .  .  .  .  . 90-91

I

Initial Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 36-37
Inserts . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 326
Insulated Wall Details . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 426-441
Insulation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 281, 445, 566
Application . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 447-450, 572
Interfacing . .  .  .  .  .  .  .  .  .  .  .  . see Joints and/or Tolerances
Interior Finishes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 239-240
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 322

J

Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 266, 364-377
Access . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 450
Architectural Treatment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 375-376
Architectural Trim Units (Cast Stone) . .  .  .  .  .  .  .  .  .  . 234
Backing Materials . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 374-375
Below Grade . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 376-377
Detailing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368-369, 375-376
Expansion . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 366-368

False . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368-369
Fire Protection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 376-377
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 364
Installation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 371, 373-375
Location . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368
Maintenance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 259
Materials . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 371-373
Number . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368
Parapet . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 541-542
Primers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 375
Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .
Sealant Depth . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 371
Seismic Seals . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 367-368
Single-Stage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 364
Staggered . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 369, 375
Stone Veneer . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 217-218
Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 359, 370
Treatment, Architectural . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 375-376
Two-Stage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 365-367
Types . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 364-367
Weathering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368
Width . .  .  .  .  .  .  .  .  .  . 317-318, 359, 362, 366, 369-370
Windows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 379-380

L

Lead Time . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 54-55, 211
LEED . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 468, 470-487
LEED Credits . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 473-487
Lettering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 35, 168-169
Life Cycle Assessment (LCA) . .  .  .  .  . 464-465, 467-468
Life Cycle Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 36, 463-464
Life Cycle Inventory (LCI) . .  .  .  .  .  .  .  .  .  .  .  .  .  . 464-468
Lifting Devices . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 291-292, 324-325
Light Shelf . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 131
Loadbearing Units
Applications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 4-6, 8, 73-80, 83-91
Connections . . 316, 319, 332-333, 344-346, 511-512
Design Considerations . .  .  .  .  .  .  . 80-83, 271, 281-282
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 296
Floor Plates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 82
Floor-to-Floor Height . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 74-75, 83
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 48, 55, 73
Hybrid Moment Resistant Frame . .  .  .  .  .  .  .  .  .  .  . 90-91
Schedule . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 75-76
Shapes and Sizes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 78-81, 282
Spandrels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 73-74, 83, 285-286
Wall-to-Floor Ratio . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 74
Loads on Precast Concrete Units. .  .  .  .  .  .  .  .  . 274-275

M

Maintenance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 259-261
Marble Expansion . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 213
Master Mold . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 46, 50, 52, 54
Masonry Tie-Back Connections . .  .  . see Connections
Matching Precast and Cast-in-Place Concrete . .  . 238-239

Materials
Attached or Incorporated . .  .  .  .  .  .  .  .  .  .  .  .  .  . 386-387
Economy of . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Specifications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 557-567
Miscellaneous Applications . .  .  .  .  .  .  .  .  .  .  .  .  .  . 24-35
Mitered Corners . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 131, 133-138
Mix Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Mockups . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 105-108, 266, 386, 555
Moisture Transfer Analysis . .  .  .  .  .  .  .  .  .  .  .  .  . 441-444
Molds . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 50-51
Lead Time . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 55
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 54, 121
Types . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 56-58
Mullions . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 68, 280-281, 297
Multiple Mixtures and Textures . .  .  . 119, 177-181, 210

N

Noise Criteria . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 491-493
Non-Loadbearing Spandrels . .  .  .  .  .  .  .  .  . see Cladding
Non-Loadbearing Walls . .  .  .  .  .  .  .  .  .  .  .  . see Cladding

O

Open Shapes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 111-112, 379-380
Other Incorporated Materials . .  .  .  .  .  .  .  .  .  .  .  .  .  . 386

P

Paint . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Panels
Arrangement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 64
		 Rotation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 316
		 Shape . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 315
Size . .  .  .  .  .  .  .  .  . 50, 58-60, 62, 78-81, 140-142, 287
Thickness . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . .289
Translation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 316
Panelization . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 58
Parapets
Copings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 540-541
Details . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 539-542
Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 539-540
Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 541-542
Patching . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 241, 325
Performance Requirements . .  .  .  .  .  .  .  .  .  .  .  . 548-550
Performance Specifications . .  .  .  .  .  .  .  .  .  .  .  . 546-547
Permeability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 254, 307, 423-425
Permeance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 423-426
Piece Size . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Panel Size
Pigments . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 145-147, 255
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 59-60
Polished and Honed . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Post-Tensioning . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 303
Pre-Bid Process . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 268-269
Conference . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 269
Samples . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 104
Precast Concrete Forms for CIP Concrete . .  .  . 94-97

ARCHITECTURAL PRECAST CONCRETE

|

585

INDEX BY SUBJECTS

Precaster Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 269-270
Prescriptive Specifications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 546
Prestressing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 289, 302-303
Production Approval Samples . .  .  .  .  .  .  .  .  .  . 104-108
Production Drawings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 299
Protection
Anchorages . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 302
Concrete Surfaces . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 256-258, 297
Hardware . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 326-327
Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 310-312
Protective Coatings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Sealers
Punched Windows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 63, 85

Q

Quality Assurance . .  .  .  .  .  .  .  . 36, 38-39, 270, 552-555
Quality Control . .  .  .  .  .  .  .  .  .  .  .  . 39, 575-576, 578-579
Quirk . .  .  .  .  .  .  .  .  .  .  .  . 58, 131, 133-134, 136-138, 140

R

Radiused Precast Concrete . .  .  .  .  .  .  .  .  .  .  .  .  . 127-129
Rain: Effects on Buildings . .  .  .  .  .  .  .  .  .  .  .  .  . 242-248
Range Samples . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 106, 554
Reglets . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 538
Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 300-312
Bends . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 305
Corrosion Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 305-312
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 61
Cover . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 184, 306-307, 310, 570
Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 270
Epoxy-Coated . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 311-312
Galvanized . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 310-311
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 300-301
Minimum . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 301
Prestressing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 302-303
Reinforcing Bars . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 302
Shadow Lines . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 304-305
Spacing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 300
Specifications . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 557-558
Tack Welding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 305
Types . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 300-303
Welded Wire Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 301
Repair . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 107, 218, 241-242, 579
Repetition . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 45, 50-54, 271-272
Responsibility . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 263-271
Architect . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 263-266
Connection Design . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 313
Construction Manager . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267-269
Engineer of Record . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 266-267
Erector . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 270-271, 292-293
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 263-264
General Contractor . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267-269
Precaster . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 265, 266, 269-270
Retarded . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Returns . .  .  .  .  .  .  .  .  .  .  .  .  . 109, 138-139, 156-157, 232
Reveals . .  .  .  .  .  .  .  .  . 53, 58-59, 113-119, 125, 368-369

586

|

ARCHITECTURAL PRECAST CONCRETE

Ribs . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 122, 289
Roofing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 536-543
Differential Wall-to-Roof Movement . .  .  .  .  .  . 538-539
Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 536-539
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 536
Scuppers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 542-543
Wall to Parapet . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 440, 539-542
Rustications . .  .  .  .  . 53, 58-59, 113-118, 125, 177, 179

S

Safing Insulation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 532-533
Samples
Assessment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100-101, 108-109
Design Reference . .  .  .  .  .  .  .  .  .  .  .  .  . 61, 551-552, 574
Development . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100, 102-103
Information Required . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 103
Mockups . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 106-108, 555
Pre-Bid . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 104
Production Approval . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 104-105
Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106, 554
Submittal . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 551-552
Weathering . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 109
Sand . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Fine Aggregate
Sand Embedment . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Sand or Abrasive Blasting . .  .  .  .  .  .  .  .  .  .  . see Finishes
Sandwich Wall Panels . .  .  .  .  .  .  . 81, 426-440, 450-459
Bowing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458-459
Fire Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 530
Insulation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 455-456
Panel Size . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 457
Thermal Bridges . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 407-410, 422
Thermal Properties . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 407-410
Wythe Connectors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 457-458
Wythe Thickness . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 456-457
Schedule
Cladding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 35-36
Loadbearing Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 75-76
Screen Units . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 396-397
Sculpture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 24
Sculpturing of Panels . .  .  .  .  .  .  .  .  .  .  .  .  . 120-124, 289
Scuppers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 542-543
Sealants . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Joints
Sealers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 256-259, 373
Seismic Design
Connection Types . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 319-320
Panel Movements . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 316-319
Seals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367-368
Shearwalls . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 81, 92-94
Self-Cleaning Concrete . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 6-7
Sequential Casting . .  .  .  .  .  .  .  .  .  . 57-58, 139-140, 177
Shading . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 394-403
Shapes
Details . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 140-141
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 78-81
In Relation to Finishes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 138-139
Open or Closed . .  .  .  .  .  .  .  .  .  .  .  .  .  . 111-112, 379-380

Shear Plates . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Connections
Shearwalls: Precast Concrete Units used as . .  . 81, 92-94
Shelf Angles . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 205
Shims . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 324, 327-328
Shipping . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Transportation
Shop Drawings . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 298-300
Approval . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 299-300
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 211
Review . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 299-300
Stone Veneer . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 211
Submittal . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 551
Shrinkage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 275
Signs . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 35
Single Source Provider . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 37
Site Access . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 267, 290-291
Sizing of Units . .  . 50, 58-60, 62, 78-81, 140-142, 287
Slab to Wall Connections . .  .  .  .  .  .  .  . see Connections
Slenderness Ratio . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 287
Smooth-Off-the-Mold . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Soffit
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Hanger Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 332, 342
Solid Wall Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 64
Smooth As-Cast . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Sound Absorption . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 489-491
Sound Barriers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 32-34
Sound Transmission Loss . .  .  .  .  .  .  .  .  .  .  .  .  .  . 488-490
Spandrel Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Cladding also
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 283-284
Deflection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 283-284
Design Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 282-285
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 66-67
Loadbearing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 73-74, 83, 285-286
Rotation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 285-286
Specification . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 545-580
Guide . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 547-580
Performance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 546-547
Prescriptive . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 546
Stacked Panel . .  .  .  .  . 26-28, 49, 69-73, 286-287, 316
Stains . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Stairs . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 32
Standoff Distance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 504-507
Stone Veneer-Faced Precast Concrete . .  . see Finishes
Story Drift . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 314, 316-319
Structural Design of Units . .  .  .  .  .  .  .  .  .  .  .  .  . 270-276
Checklist . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 273-274
Design Criteria . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 272-273
Design Objectives . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 271-272
Determination of Loads . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 274-275
General Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 271
Volume changes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 275-276
Structural Steel Framing Tolerances . .  .  .  .  .  .  .  .  . 357
Submittals . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 550-552
Sunscreens . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 32
Surface Aesthetics . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 35, 36, 99

Surface Condensation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 420-423
Surface Out-of Planeness . .  .  .  .  .  .  .  .  .  .  .  .  .  . 349-350
Sustainability . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 37, 459-487
Concepts . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 460-463
Cost . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 461-464
Definitions . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 459-460
LEED Credits . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 469, 473-487
Life Cycle Analysis . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 463-468
Production of Precast Concrete . .  .  .  .  .  .  .  .  .  . 478-482
Rating Systems . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 469-473

T

Tack Welding . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 305
Tendon Anchorages, Protection . .  .  .  .  .  .  .  .  .  .  .  .  . 302
Terra Cotta . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Texture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see also Finishes
Thermal Bridges . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 407-410, 422
Thermal Expansion, Coefficient of . .  .  .  .  .  .  .  .  .  . 275
Thermal Mass . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 410, 412-417
Thermal Resistance (R-value) . .  .  .  .  .  .  . 402, 404-410
Thickness
Facing Mixture . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 110, 571
Fire Resistance, Relation to . .  .  .  .  .  .  .  .  .  .  .  .  . 527-532
Panel . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 287, 289
Sandwich Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 454, 530-532
Wythe . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 456-457
Tie-Back (Lateral) Connections . .  .  .  . see Connections
Thin Brick . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 190-191, 202-203
Standard . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 189
Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 347-364
Acceptability of Appearance . .  .  .  .  .  .  .  .  .  .  .  . 240-241
Anchor Bolts . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 358
Architectural Trim Units (Cost Stone) . .  .  .  .  .  . 233-235
Building Frame . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 356-357
Bowing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348-350
Clearances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 362-363
Concrete Requirements . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 110
Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 320-321, 357
Erection . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 235, 350, 356-363, 578
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 347
Interfacing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 363-364
Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 359, 370
Local Smoothness . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 349-350
Primary Control Surface . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 356
Product . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 347-355, 572-574
Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 307
Secondary Control Surface . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 356
Warping . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 348-350
Windows . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 112
Tooling . .  .  .  .  .  .  .  .  .  .  .  .  . see Bushhammering (Finishes)
Top Floor: Usage for . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 47
Torsion . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 283-284
Towers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 28-31
Transportation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 140-144, 556
Two-Stage Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 365-367
Two-Stage Precasting . .  .  .  .  .  .  .  . 57-58, 139-140, 177

ARCHITECTURAL PRECAST CONCRETE

|

587

INDEX BY SUBJECTS

U

Uniformity of Appearance. . 59-60, 99-102, 109, 153, 157,
169, 174, 182, 184, 211
Urban Heat Island . .  .  .  .  .  .  .  .  .  . 37, 393-394, 475-477
Uses of Precast Concrete: see Applications
U-values . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 405, 407

V

Vapor Retarders . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 423-444
Veneer-Faced Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Finishes
Volume Changes . .  .  .  .  .  .  .  .  .  .  .  .  .  . 275-276, 313-314

W

Wall Panel Systems . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 62
Wall-Supporting Units . .  .  .  .  .  .  .  .  .  . 26-28, 49, 69-73
Wall-to-Floor Ratio . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 74-75
Wall to Foundation Connections . .  . see Connections
Wall to Wall Connections . .  .  .  .  .  .  .  . see Connections
Warping, Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 348-350
Water Absorption . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Absorption
Water-Cement Ratio . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 100
Water Drips . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Drips
Water Leakage . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 441
Weathering . .  .  .  .  .  .  .  .  .  .  .  .  .  . 212, 242-250, 255-256
Absorption . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 245, 254-256, 567
Aggregate, Effect on . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 148-149
Causes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 242-248
Concrete Design for . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 255-256
Efflorescence . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 250-255
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 242-248
Glass Staining . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 387-390
Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 368-369
Sealers . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 256-258
Surface Finish . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 248-250
Water Drips . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . see Drips
Water Run-off . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 242-250
Waterfall . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 34-35
Weep Holes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 206
Welded Wire Reinforcement . .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 301
Welded Connections . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 328-329
Welding
Protection During . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 297
Stainless Steel . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 327
Windows
Acoustical Properties . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 493-496
Blast Resistance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 520-523
Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . 259, 390
Condensation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 447
Design Considerations . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 379-384
Detailing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 382-384
Exposed Aggregate . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 380
Fire Endurance . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 530-532
Flashing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 381

588

|

ARCHITECTURAL PRECAST CONCRETE

Frames . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 382, 386, 459, 521-522
General . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 379
Glazing . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 393, 493-496, 520-523
Installation . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 384-386
Joints . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 379-380
Mockups . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 386
Open and Closed Shape Panels . .  .  .  .  .  . 111-112, 382
Punched . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 63, 85
Sealant . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 380
Shading . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 394-403
Sills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382-383
Staining or Etching . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 387-390
Tolerances . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 379
Washing Systems . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 386-387
Weep Holes . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 381-382
Window Wall Panels . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 65-66
Wythe Connectors . .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  .  . 457-458



Source Exif Data:
File Type                       : PDF
File Type Extension             : pdf
MIME Type                       : application/pdf
PDF Version                     : 1.6
Linearized                      : Yes
Encryption                      : Standard V2.3 (128-bit)
User Access                     : Print, Copy, Extract
Page Mode                       : UseOutlines
XMP Toolkit                     : Adobe XMP Core 4.0-c316 44.253921, Sun Oct 01 2006 17:08:23
Instance ID                     : uuid:4546c795-a1e8-4514-9350-d9f48f1a6172
Document ID                     : adobe:docid:indd:6b4493cd-0f32-11db-8d6b-843ee7574fee
Rendition Class                 : proof:pdf
Manifest Link Form              : ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream, ReferenceStream
Manifest Placed X Resolution    : 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00
Manifest Placed Y Resolution    : 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00, 72.00
Manifest Placed Resolution Unit : Inches, Inches, Inches, Inches, Inches, Inches, Inches, Inches, Inches, Inches, Inches
Manifest Reference Instance ID  : uuid:5142CE91403311DCAC5BDA4C0EFA6957, uuid:fb7f7dda-0df1-11db-8a4b-000d9342bde0, uuid:31c334b7-a526-490d-8de6-ec1d2496f1a5, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0, uuid:8d600027-0deb-11db-8a4b-000d9342bde0
Manifest Reference Document ID  : uuid:5142CE90403311DCAC5BDA4C0EFA6957, uuid:F5E9BCA90F5311DBABF5F8094E8A1C17, uuid:A87ADA31469411DC9F5DD0EB24E5B30C, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17, uuid:F5E9BCA70F5311DBABF5F8094E8A1C17
Create Date                     : 2007:08:14 11:16:09-05:00
Modify Date                     : 2007:08:16 15:59:22-05:00
Metadata Date                   : 2007:08:16 15:59:22-05:00
Creator Tool                    : Adobe InDesign CS3 (5.0)
Thumbnail Format                : JPEG
Thumbnail Width                 : 256
Thumbnail Height                : 256
Thumbnail Image                 : (Binary data 3831 bytes, use -b option to extract)
Format                          : application/pdf
Title                           : PCI Architectural Precast Concrete Manual (MNL-122)
Creator                         : Sidney Freedman
Rights                          : ©2007 Precast/Prestressed Concrete Institute
Authors Position                : Director of Architecural Services
Marked                          : True
Startup Profile                 : Print
Producer                        : Adobe PDF Library 8.0
Trapped                         : False
Has XFA                         : No
Page Count                      : 609
Page Layout                     : SinglePage
Author                          : Sidney Freedman
EXIF Metadata provided by EXIF.tools

Navigation menu