Genetic Characterization Of Liriodendron Seed Orchards With EST SSR Markers LT157 2050 2389 4 1

User Manual: LT157

Open the PDF directly: View PDF PDF.
Page Count: 9

DownloadGenetic Characterization Of Liriodendron Seed Orchards With EST-SSR Markers LT157 2050-2389-4-1
Open PDF In BrowserView PDF
Journal of Plant Science & Molecular Breeding
ISSN 2050-2389 | Volume 4 | Article 1

Research				

Open Access

Genetic characterization of Liriodendron seed orchards
with EST-SSR markers
Xinfu Zhang1, Alanna Carlson1, Zhenkun Tian2, Margaret Staton3, Scott E. Schlarbaum4, John E. Carlson5 and Haiying Liang1*
*Correspondence: hliang@clemson.edu

CrossMark
← Click for updates

Department of Genetics and Biochemistry, Clemson University, Clemson SC 29634, USA.
Beijing Forestry University, Beijing, China.
3
Department of Entomology and Plant Pathology, The University of Tennessee, Knoxville, TN 37996-4563, USA.
4
Department of Forestry, Wildlife & Fisheries, The University of Tennessee, Knoxville, TN 37996-4563, USA.
5
Department of Ecosystem Science and Management and The Department of Plant Science, Pennsylvania State University,
University Park, PA 16802, USA.
1
2

Abstract
Liriodendron tulipifera L., is a wide-spread, fast-growing pioneering tree species native to eastern North
America. Commonly known as yellow-poplar, tulip tree, or tulip-poplar, the species is valued, both
ecologically and economically. It is perhaps the most commonly used utility hardwood in the USA, and
is planted widely for reforestation and, in varietal forms, as an ornamental. Although most seedlings
used for reforestation today derive from collections in natural populations, two known seed orchards,
established from plus-tree selections, i.e. superior phenotypes, in the 1960’s and 1970’s have been used for
local and regional planting needs in Tennessee and South Carolina. However, very little is known about
the population genetics of yellow-poplar nor the genetic composition of the existing seed orchards. In
this study, 194 grafted yellow-poplar trees from a Clemson, SC orchard and a Knoxville, TN orchard
were genetically characterized with 15 simple sequence repeat (SSR) markers developed from expressed
sequence tags (ESTs). Of the 15 EST-SSR markers, 14 had a polymorphic information content (PIC) of
at least 0.5. There was no significant difference between the Clemson and Knoxville orchards in average
effective number of alleles (5.93 vs 3.95), observed and expected heterozygosity (Ho: 0.64 vs 0.58; He: 0.74
vs 0.70), Nei’s expected heterozygosity (0.74 vs 0.58), or Shannon’s Information index (1.84 vs 1.51). The
larger Clemson orchard exhibited a significantly greater number of observed alleles than the Knoxville
orchard (15.3 vs7.4). Overall, substantial genetic diversity is captured in the Clemson and Knoxville
orchards.
Keywords: Genetic diversity, seed orchard, SSR markers, species

Introduction

Liriodendron tulipifera L., commonly known as yellow-poplar
or tulip-poplar, is a wide-spread, fast-growing pioneering
hardwood species of considerable economic value in the
forests of eastern North America. Yellow-poplar is distributed
predominantly east of the Mississippi River from the gulf coast
to southern Canada (28° to 43° north latitude) [35]. According
to the forest inventory analysis [11], as surveyed from 20062012, the total saw log volume of L. tulipifera on timberland in
the United States was 25.9 billion cubic feet, with the majority
(65%) located in the southeastern United States. The species is
shade intolerant and highly competitive, growing faster than

Acer rubrum L. (red maple) and Quercus rubra L. (northern red
oak) seedlings under a variety of silvicultural understory treatments (Beckage and Clark 2003). Yellow-poplar is often seen as
a pioneering species in old fields. As a component of 16 forest
cover types, this species’ degree of dominance has created
differentiation between the ecological communities [46]. In
addition, yellow-poplar is valued as a nectar source for honey
production, as a source of wildlife food (mast), and as a large
shade tree in urban plantings [3]. The wood of yellow-poplar is
used in a diverse range of products, such as in furniture, pallets
and framing construction as well as pulp [12,41]). Chemical
extracts from yellow-poplar wood or leaves have proven useful,

© 2015 Liang et al; licensee Herbert Publications Ltd. This is an Open Access article distributed under the terms of Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0). This permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

such as sesquiterpenes which have an anti-tumor effect and
antifeeding for herbivores [27], and antimicrobial alkaloids [2].
L. tulipifera has been cultivated since 1663 [5] and is currently
widely planted in eastern forests. Although seed orchards have
been established to meet local or regional planting needs
in the U.S.A. [6,36], genetic diversity of Liriodendron seed
orchards in relation to natural stands has not been studied.
Because seed orchards is the bridge between breeding and
silvicultural activities, genetic diversity of tree seeds orchards
determines the genetic quality of future forest stands and
forms the basis for further improving the management of
genetic resources and for the genetic modification of cultivars
to meet new environmental challenges. Thus, the lacking
information limits utilization of these Liriodendron orchards
in a tree improvement program.
The primary goal of our study was to determine the genetic
composition and diversity in two Liriodendron seed orchards
in the southeastern USA. The orchard residing in Knoxville,
Tennessee, was established in 1966 and contains 100 grafted
ramets, representing 31 genotypes or clones. The Clemson
orchard in South Carolina was established in 1976 by grafting
multiple ramets of 150 plus trees selected from throughout
the 17,500-acre Clemson Experimental Forest by Dr. Roland E.
Schoenike (http://www.clemson.edu/trails/history/schoenike.
html#top). Seeds from this orchard have been used for
reforestation efforts for a number of years. Currently there
are 165 surviving trees in the Clemson orchard. Besides L.
tulipifera, the only other Liriodendron species is Liriodendron
chinense, which is native to China and Vietnam.
Although the two species separated 10~16 million years
ago [32], they are quite similar morphologically and are
cross fertile [26,34], and the hybrids exhibit heterosis [31,39].
Because the incomplete records suggest that the Clemson
orchard may contain L. chinense or hybrids, we first used
the sequence of a chloroplast gene, maturase K (matK), to
discriminate the two Liriodendron species and their hybrids.
Then we investigated the genetic diversity and allele richness
among selections of this unique native species in each orchard
as a first step toward contrasting orchard-produced seedling
diversity with natural diversity. We chose simple sequence
repeat (SSR) markers (also called microsatellites) in the study,
because SSR markers are co-dominant, easily reproduced and
scored, highly polymorphic, abundant through the genome,
and have higher information content than isoenzyme and
dominant markers [45].

Materials and methods

Plant materials and DNA isolation

Fresh leaves of all Liriodendron trees (165) from the Clemson
seed orchard and 31 trees from the Knoxville seed orchard
were collected in the spring of 2013 and stored in plastic bags
at -80°C prior to DNA isolation. All these trees represented
different clones as validated by the SSR markers used in this
study. Leaves from a Liriodendron tulipifera tree (accession

doi: 10.7243/2050-2389-4-1
number 70921 H) from the US National Arboretum (collected
by Kevin Conrad) were also included in the study. Total
genomic DNA was isolated from leaves using a CTAB protocol
as described in [16] and suspended in TE buffer (Tris base
6.1g/L, EDTA 0.37 g/L, pH 8). The quality and concentrations
of genomic DNA from individual plants were determined with
a NanoDrop 3300 (Thermo Scientific, Wilmington, Delaware,
USA) and by electrophoresis on 0.8% agarose gels.

Distinguishing between two Liriodendron species based
on maturase K sequence

The record of the 165 surviving Liriodendron trees in the
Clemson orchard is not complete. Therefore, the sequence of a
chloroplast gene, maturase K (matK) was used to discriminate
between the species/hybrids. The matK sequence was amplified
with forward (5’-CGATCTATTCATTCAATATTTC-3’) and reverse
primers (5’-TCTAGCACACGAAAGTCGAAGT-3’) in a 12.5-μl
reaction containing 6.875 uL ddH2O, 1 uL MgCl2 (25 mM),
0.5 uL forward primer (10uM), 0.5 uL reverse primer (10uM),
0.25 uL dNTPs (10 mM each), 0.25 uL BSA (0.8ug/uL), 0.125
uL Taq Pololymerase (5u/uL), 0.5 uL DNA (~20ng/ul), 2.50 uL
5X PCR buffer (-Mg).
The conditions for polymerase chain reactions (PCR) were
as follows: 5 minutes of initial denaturation at 94°C, 35 cycles
of touch-down PCR with 30 seconds of denaturation at 94°C,
30 seconds of annealing at 60-50°C (first cycle 60°, then each
subsequent cycle 1°C lower than the previous until 51°C
annealing temperature, followed by 25 cycles each with a
50°C annealing temperature), and 3 minutes of extension
at 72°C, and a final extension at 72°C for 10 minutes. Before
being sequenced with 1 ul of 10 uM forward or reverse
primer, PCR products were cleaned with ExoAP mix (89 uL
H2O+ 10 uL 5000U/mL Antarctic Phosphatase +1 uL 20000U/
mL Exonuclease I) for 30 minutes in a reaction containing 1
uL of PCR product and 1uL of ExoAP mix, followed by a heat
inactivation step at 80°C for 15min. An 834 bp-segment of
maturase K gene from each tree was used for alignment with
MUSCLE and curated with Gblocks, and a phylogenetic tree
was built with maximum likelihood (PhyML) (http://www.
phylogeny.fr/) [7].
The maturase K gene sequence of L. tulipifera (GI: 5731451), L.
chinense (GI: 7239759), and a hybrid (GI: 389955358) available
in GenBank were included in the analysis.

L. tulipifera EST-SSR markers, PCR amplification, and
allele sizing

Twenty simple sequence repeat (SSR) markers (also called
microsatellites) were used to investigate the genetic
composition of the Liriodendron seed orchards. These
markers included seven Expressed Sequenced Tags (EST)SSR markers (LT002, LT015, LT021, LT086, LT096, LT131,
LT157) previously characterized by electrophoresis on 8%
polyacrylamide gels [42] and thirteen new markers (LTCU19,
LTCU40, LTCU51, LTCU53, LTCU125, LTCU139, LTCU142, LTCU143,

2

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

LTCU145, LTCU150, LTCU151, LTCU152, LTCU154) mined from
a comprehensive EST dataset [22]. PCR amplification for each
marker was performed with genomic DNA of Liriodendron
trees from the Clemson and Knoxville seed orchards and the
US National Arboretum. For a more cost-effective 153 primer
screening, a M13 tail (5’-CACGACGTTGTAAAACGAC-3’) was
added to the 5’-end of the forward primer of each marker
pair in order to amplify the fragments using a complementary
adapter with a fluorescent dye (6-FAM, VIC, NED, or PET) at
its 5’-end (Applied Biosystems, Foster City, California, USA).
Polymerase chain reactions were carried out in a 12.5-μl
solution comprising: approximate 75 ng DNA template,
0.052 U/μL Promega Taq DNA polymerase, 0.16 nM forward
primer, 0.4 nM reverse primer, 0.4 nM fluorescent M13 primer,
0.24 mM each dNTPs, and 1.2×Promega PCR buffer. The
PCR profile consisted of an initial denaturation at 94°C for 3
minutes followed by 10 cycles of 1 minute at 94°C, 1 minute
at annealing temperature (Table 1), and 1 minute 15 seconds
at 72°C, and then 35 cycles of 1 minute at 94°C, 1 minute at
58°C, and 1 minute at 72°C, with a final extension of amplified
DNA at 72°C for 5 minutes.
An aliquot of 1.5 μl PCR products were treated with 1.5 μl
of 10-fold diluted ExoSAP-IT (Affymetrix Inc. Cleveland, OH,
USA) to remove single stranded primers which might influence
fragment analysis at 37°C for 30 minutes and then at 80°C
for 15 minutes. After being diluted to 100 ng/μl, 1μl of each
sample was mixed with 0.1 μl of LIZ600 and 8.9 μl of Hi-Di
Formamide, denatured at 95°C for 5 minutes, and then put
on ice for 10 minutes before being separated on an ABI 3730
Genetic Analyzer. The Dye set was DS-33 (6-FAM, VIC, NED,
PET and LIZ). Allele sizes were scored with GeneMapper (4.0)
(Applied Biosystems, Foster City, California, USA). Functional
annotation of EST-SSRs was performed by applying a homology
search of reassembled ESTs against the non-redundant (nr)
NCBI database using the BLASTx algorithm [1].

Data analysis

MICRO-CHECKER [38] was employed to check for potential
genotyping errors arising from large allele drop-out and
stuttering. Observed and expected heterozygosities and
polymorphic information content (PIC) were calculated using
Cervus 2.0 [25]. Deviations from Hardy–Weinberg equilibrium
and the Shannon’s Information index were calculated with
GENEPOP (http://genepop.curtin.edu.au/, Raymond and
Rousset 1995).

doi: 10.7243/2050-2389-4-1
lobed leaves and smaller flowers. However, our attempt to
tell these two species apart by morphology failed: the leaf
shape varied depending on age (Supplementary figure S1)
and the flowers were located at a too high for sampling.
Molecular techniques including biochemical analysis [34],
isozymes [14], and fingerprinting with random amplified
polymorphic DNA (RAPD) [21] have been explored in
discrimination of Liriodendron species and their hybrids. In
2012, Zhang et al., reported an SSR marker that amplified a
190-bp fragment from L. chinense, a 180-bp fragment from L.
tulipifera, and both 190- and 180-bp fragments from hybrid.
In this study, matK sequence was employed. The matK gene
locates within the intron of the trnK and codes for maturase
like protein involved in Group II intron splicing [37]. The
trnKUUU-matK region, ranging from approximately 2.2 kb
(liverworts) to 2.6 kb (seed plants) in size, is universally present
in land plants and only few exceptions of a secondary loss
or reorganizations are known to date [40]. Because the matK
gene evolves more rapidly, compared to other plastid genes,
it has become a valuable marker for lower-level phylogenetic
reconstruction of systematic and evolutionary studies. The
Clemson seed orchard contains 165 surviving trees. The
matK sequence was amplified from each of the 165 trees in
the Clemson orchard (Supplementary figure S2). When the
amplicons were pair-end sequenced, an 834-bp segment of
high quality was obtained for each tree, representing 55%
of the full-length gene. There were only eight nucleotides
different between the two Liriodendron species within the 834bp segment. As shown in Figure 1 and Supplementary figure S3,
only Tree#CU24 and #134 were not L. tulipifera. Their hybrid
status was established by having seven nucleotides different
from L. tulipifera and 2 nucleotides different from L. Chinense.
These results confirm the record of hybrids being planted in the
Clemson Orchard. Thus, these two hybrids were excluded in
the genetic composition analysis. Further, our study indicates
that L. tulipifera, L. Chinense, and their hybrids contain unique
nucleotide compositions in matK sequence that can be utilized
in distinguishing the species and hybrids.

Amplification of EST-SSR loci in Liriodendron

No evidence for large allele dropout was found for any of the
20 markers. Stuttering occurred in five markers: LT131, LT157,
LTCU40, LTCU142, and LTCU143, and these five markers were
excluded in further analyses. All of the remaining 15 markers
were polymorphic in both Clemson and Knoxville orchards.
The 20 markers were also tested on one L. tulipifera and one
Results and discussion
L. chinense tree from the US National Arboretum. Eleven loci
Distinguishing between two Liriodendron species with were heterozygous and one locus failed in the L. tulipifera
maturase K (matK) sequence
tree. PCR amplification for all 20 markers was successful in
It is not clear when L. chinense was first introduced to U.S.A., the L. Chinense tree, although sizing in an ABI 3730 Genetic
but the two Liriodendron species hybridize readily [26] and Analyzer failed for LT157 and LTCU142, due to stuttering.
efforts in crossing have been well-documented [e.g., 31,34]. Fourteen loci were heterozygous in the L. chinense tree. This
The two species are similar morphologically, except that L. indicates a high frequency of transferability of L. tulipifera
chinense is smaller in stature and has larger, more deeply EST-SSR markers in L. Chinense, supporting the previous

3

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

doi: 10.7243/2050-2389-4-1

Table 1. Characteristics of 20 EST-SSR loci.
Marker
name

Repeat motif

Forward primer
sequence (5’-3’)

LT002

(GCA)8

CCTACCACCAGCA

TCTCGTCGCTGAAGAT

ATACCTA

ATG

LT015

(CCGAAC)5

LT021

(TTC)8

LT086

(CTT)10

LT096

(CT)20

LT131

(AC)22

LT157

(TTC)6

LTCU19

(AG)10

LTCU40
LTCU51

(ATG)8
(CT)18

LTCU53

(TG)14

LTCU125

(TC)8

LTCU139

(TCT)10

LTCU142

(AAT)8

LTCU143

(TG)13

LTCU145

(GA)18

LTCU150

(TC)10

LTCU151

(TC)11

LTCU152

(CA)17

LTCU154

(CT)10

Reverse primer
sequence (5’-3’)

Expected size Stuteringa Annealing
(bp)
temperature ⁰C
189

N

59

110

N

59

180

N

57

274

N

55

272

N

55

TCCGTTATCTCTCT

CTAGACAGGTGCTCGG

CCAAAA

ATAC

CAAATACCATTGC

ACGCATCCTCTTCCAC

ACCTTGT

TAC

AAGACAGGACTTT

GAACGAACCTAACCA

CCACTGA

AATGA

TGCAACCTAACAA

TGAAAAGCAACCAAG

GATGTGT

TTACC

GCAGCATCTCCTC

TTGCAGTTGAGCTATT

240

Y

55

ATATTCT

GTTG

AGTTGCCCTTTAGC

GCCACAGAGTTTTGGA 222

Y

55

TTCTTT

AGTA

GTGGATTGCAAAG

AAAACAAAAGCAAGC

183

N

57

GCAGAGT

AAGCC

TTGCGTAAATGCA

GAAGCCtaTGCAAGAT

181

Y

55

TCCAAAA

GCAA

ATCACCATCTTCCT

AAACCATTCCAACCAT 198

N

55

CATCGC

CCAA

CGGATCTTTCTCTT

AAGAAGATTGCAGAG

223

N

55

TCCATCC

GCAGAA

CGAAAGACATTCC

CCATTACAATCCACAG

205

N

55

CATCACA

CCAA
164

Y

55

171

Y

55

160

N

55

157

N

55

167

N

55

189

N

55

177

N

55

156

N

55

GAATAACCGCTCT

AAGCCAAGTGGCAAA

TTTGGGA

GAAGA

TGGTGCATATGGG

TATTCCCCCAGCTTCT

CTTAGAA

CCTT

AAAAATGCTAATC

TATCCAACCGATCACC

CAATAACTTTCG

CATT

TTGAAGTCCAGAT

GCCTAGGGaGATGtTTT

TGATTGATTG

TGG

TCTTCAAACCAAG

GCACTACATCCCTTTTc

GCTGTTG

CCA

TGAGGTGACTTTG

GACCCgaGCTGTAAAA

GCTTTTG

TGGA

CATCCAAATGCAG

ATTCCCACTCGGTTGA

CAGAAAT

ACAC

GATGAAGGAGAAT

CCAGCCAAGAAAGAA

TCTATATTTTCTGA

AATGG

Y:Yes; N:No.

a

4

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

1

1

0.003

Chinese_GI_7239759
Hybrid_GI_389955358
CU24
CU134
CU16
Tul_GI_5731451
Tul_ACC70921H
CU2
CU3
CU4
CU5
CU6
CU7
CU8
CU9
CU10
CU13
CU14
CU17
CU18
CU19
CU21
CU22
CU23
CU75
CU76
CU78
CU96
CU118
CU119
CU124

Figure 1. Sequence comparison of maturase K gene. The
sequences were aligned with MUSCLE the phylogenetic tree
was built with maximum likelihood (PhyML). The maturase K
gene sequence of L. tulipifera (GI: 5731451), L. chinense (GI:
7239759), and a hybrid (GI: 389955358) available in GenBank
were included in the analysis. The complete illustration of the
tree is included in the Supplementary figure S1.

doi: 10.7243/2050-2389-4-1
Table 2. Statistics of the 15 markers analyzed by cervus.
Locus

K

N

HO

HE

LT002

6

174

0.72

0.69 0.65

PIC

LT015

6

172

0.55

0.60 0.54

LT021

3

173

0.17

0.19 0.18

LT086

7

170

0.36

0.53 0.50

LT096

15

165

0.65

0.76 0.73

LTCU19

12

174

0.60

0.71 0.69

LTCU51

18

172

0.74

0.87 0.85

LTCU53

13

167

0.63

0.84 0.82

LTCU125

18

164

0.88

0.89 0.88

LTCU143

14

164

0.74

0.80 0.76

LTCU145

11

172

0.84

0.86 0.84

LTCU150

15

172

0.51

0.74 0.72

LTCU151

11

158

0.51

0.76 0.73

LTCU152

19

143

0.65

0.93 0.92

LTCU154

26

160

0.74

0.93 0.92

Average

13

167

0.62

0.74 0.71

K: number of alleles; N: number of individuals; H0: observed
heterozygosity; He: expected heterozygosity; PIC: polymorphic
information content.

Genetic composition of the L. tulipifera orchards

While there were only two loci (LT002 and LT015) significantly
deviating from Hardy-Weinberg proportions in the Clemson
population, there were 10 deviating loci in the Knoxville
findings of 72.4% success rate by [42] and 82.1% by [43]. This population (p>0.05) (Supplementary Table S1 and S2). This
is expected because EST-SSRs have generally demonstrated may be due to insufficient sample size from the Knoxville
a high frequency of cross-species transferability despite less population. As shown in Tables 3 and 4, the Clemson orchard
polymorphism compared to genomic SSRs [9,12,44]. Among had higher values for observed number of alleles (15.3 vs
the 194 L. tulipifera trees included in the study, the number 7.4), effective number of alleles (5.9 vs 4.0), observed (Ho,
of alleles per locus ranged from 3 to 26 (mean=13.0) (Table 2). 0.64 vs 0.58) and expected heterozygosity (He, 0.74 vs 0.70),
The observed and expected heterozygosities (Ho and He) Nei’s expected heterozygosity (0.74 vs 0.58), and Shannon’s
ranged from 0.17 to 0.89 and from 0.19 to 0.93, with averages Information index (1.85 vs 1.51) than the Knoxville orchard.
of 0.62 and 0.74, respectively. The polymorphic information However, the differences were not statistically significant
content (PIC) ranged from 0.17 to 0.92, with an average of (p=0.05, t-Test) except for observed number of alleles. The
0.71. Overall, 14 of the 15 markers had a PIC≥0.5.
different number of trees from the orchards included in the
Many genomic resources, such as expressed sequence study, 163 from the Clemson vs 31 from the Knoxville, may
tag (EST) databases [15,22,23] and genomic DNA libraries be a contributing factor.
[24], have been developed for L. tulipifera. Through these
This is the first report of genetic composition of Liriodendron
resources, several thousand putative SSR markers have been cultivated populations in North America and has provided
identified by in silico mining. However, only 345 L. tulipifera the basic data of genetic diversity and allele richness among
SSR markers having been tested for polymorphism by selections of this unique native species. [43] examined 27
polyacrylamide denaturing gels [42,43]. Compared to other trees from a cultivated population of L. tulipifera in the Jurong,
species, Liriodendron has lacked development of polymorphic Jiangsu Province of China with 39 polymorphic EST-SSR loci
and informative SSR markers.
through electrophoreses in 6% polyacrylamide denaturing
As a result, no genetic linkage maps of Liriodendron have gels and visualization with silver nitrate staining. It was found
been reported. This is in contrast with the species’ecological that the number of alleles per locus ranged from three to 18
and economic value and phylogenetic position as a basal and the average Ho and He were 0.68 and 0.78, respectively.
angiosperm.
Compared to this cultivated population in China, the two

5

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

doi: 10.7243/2050-2389-4-1

Table 3. Genetic variation at 15 EST-SSR loci characterized in the clemson orchard.
Clemson orchard (163 trees)
Sample size
326
324
324
318
308
326
324
310
308
306
324
322
294
258
302
312
--

Locus
LT002
LT015
LT021
LT086
LT096
LTCU19
LTCU51
LTCU53
LTCU125
LTCU143
LTUCU145
LTUCU150
LTUCU151
LTUCU152
LTUCU154
Mean
St. Dev.

Na
6.00
8.00
6.00
9.00
18.00
15.00
18.00
13.00
26.00
14.00
12.00
14.00
9.00
25.00
37.00
15.33
8.53

Ne
3.16
2.30
1.99
1.84
3.75
3.20
7.74
4.84
11.46
4.84
7.49
2.90
2.99
15.89
14.49
5.93
4.59

Obs_Hom
0.28
0.46
0.54
0.69
0.38
0.40
0.26
0.30
0.05
0.23
0.17
0.42
0.54
0.33
0.28
0.36
0.16

Obs_Het
0.72
0.54
0.46
0.31
0.62
0.60
0.74
0.70
0.95
0.77
0.83
0.58
0.46
0.67
0.72
0.64
0.16

Exp_Hom
0.31
0.43
0.50
0.54
0.26
0.31
0.13
0.20
0.08
0.20
0.13
0.34
0.33
0.06
0.07
0.26
0.15

Exp_Het
0.69
0.57
0.50
0.46
0.74
0.69
0.87
0.80
0.92
0.80
0.87
0.66
0.67
0.94
0.93
0.74
0.15

Nei’s
0.68
0.57
0.50
0.46
0.73
0.69
0.87
0.79
0.91
0.79
0.87
0.66
0.67
0.94
0.93
0.74
0.15

I
1.35
1.11
1.01
1.00
1.80
1.71
2.26
1.89
2.75
1.81
2.18
1.61
1.38
2.91
3.03
1.85
0.66

Na: Observed number of alleles. Ne: Effective number of alleles (Kimura and Crow 1964). Obs_Hom/Obs_
Het: Observed homozygosity/heterozygosity. Ext_Het/Exp_Het: expected homozygosity/heterozygosity
(Levene 1949). Nei's (1973) expected heterozygosity. I=Shannon's Information index (Lewontin 1972). St.
Dev.: Standard deviation.
Table 4. Genetic variation at 15 EST-SSR loci characterized in the knoxville orchard.
Locus

Knoxville orchard (31 trees)
Sample size Na Ne Obs_Hom Obs_Het Exp_Hom Exp_Het Nei’s

I

LT002

62

5

3.65 0.32

0.68

0.26

0.74

0.73

1.43

LT015

60

5

3.38 0.43

0.57

0.28

0.72

0.7

1.36

LT021

62

2

1.17 0.84

0.16

0.85

0.15

0.15

0.28

LT086

62

6

2.81 0.35

0.65

0.35

0.65

0.64

1.22

LT096

62

10

4.75 0.32

0.68

0.2

0.8

0.79

1.85

LTCU19

62

9

3.59 0.52

0.48

0.27

0.73

0.72

1.68

LTCU51

60

12

5.84 0.27

0.73

0.16

0.84

0.83

2

LTCU53

62

6

2.77 0.74

0.26

0.35

0.65

0.64

1.32

LTCU125

60

15

6.14 0.3

0.7

0.15

0.85

0.84

2.17

LTCU143

60

8

5.26 0.33

0.67

0.18

0.82

0.81

1.8

LTUCU145

60

8

4.64 0.1

0.9

0.2

0.8

0.78

1.71

LTUCU150

62

5

3.29 0.68

0.32

0.29

0.71

0.7

1.33

LTUCU151

56

3

2.26 0.39

0.61

0.43

0.57

0.56

0.89

LTUCU152

62

9

5.88 0.42

0.58

0.16

0.84

0.83

1.93

LTUCU154

58

8

3.83 0.24

0.76

0.25

0.75

0.74

1.62

Mean

61

7.4 3.95 0.42

0.58

0.29

0.70

0.58

1.51

St. Dev.

--

3.4 1.44 0.20

0.20

0.17

0.17

0.16

0.46

Na: Observed number of alleles. Ne: Effective number of alleles (Kimura and Crow 1964). Obs_
Hom/Obs_Het: Observed homozygosity/heterozygosity. Ext_Het/Exp_Het: expected homozygosity/
heterozygosity (Levene 1949). Nei's (1973) expected heterozygosity. I=Shannon's Information index
(Lewontin 1972). St. Dev.: Standard deviation.

6

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

doi: 10.7243/2050-2389-4-1

US orchards had slightly lower values of average Ho and
L.tulipifera in Clemson orchard
He, with 0.64 (Ho) and 0.74 (He) in the Clemson orchard and
L.tulipifera in Knoxville orchard L.tulipifera
L.tulipifera in National Arboretum
0.58 (Ho) and 0.70 (He) in the Knoxville orchard. These values
Hybrids in Clemson orchard
L.tulipifera X chinense
are comparable to those reported in a L. chinense cultivated
L.chinense
in
National
Arboretum
L.chinense
population in China, which had a Ho and He of 0.48 and 0.74
[43]. Other forest tree species have similar heterozygosities
0
50 40 30 20 10
as well, for example, a Pinus merkusii parental and seedling
Figure
2.
The
UPGMA
dendrogram based on Nei's (1978)
populations had a He of 0.55 and 0.49, respectively [10,30].
genetic
distance.
Bootstrap
replicates=1,000.
Reported 0.48 (Ho) and 0.63 (He) in a white spruce plantation
and 0.49 (Ho) and 0.63 (He) in a white spruce improvement
selection population. It is noteworthy that genetic diversity of and provides a foundation for further genetic and breeding
natural L. tulipifera populations has been reported [e.g., 18,32]. exploration. The polymorphic markers developed in this
However these studies utilized either allozymes or amplified study will serve as a resource enabling the future study of
fragment length polymorphism (AFLP) markers, which usually population dynamics and adaptive variation in Liriodendron.
have lower information content than SSR markers. None of
the reported expected heterozygosities from these studies Additional files
exceeded 0.29. Overall, substantial genetic diversity is captured
Supplementary Table S1
in the Clemson and Knoxville seed orchards.

Conclusion

The data obtained in this study will be useful in future
applications such as prediction of genetic gain and gene
diversity in the seed orchards. Nei’s genetic distance between
the two orchards was 0.39, which was the lowest among all
comparisons (Table 5). The L. chinense and L. tulipifera trees
from the National Arboretum exhibited the largest genetic
distance (1.17). The two orchards and the L. tulipifera sample
from the US National Arboretum grouped together in the
UPGMA dendrogram. The genetic distance of the hybrids in
the Clemson orchard was closest to the Clemson orchard (0.50),
followed by the Knoxville orchard (0.80) and L. chinense from
the National Arboretum (0.88), and then by the L. tulipifera from
the National Arboretum (1.17) (Figure 2). With a widespread
range of distribution, L. tulipifera has adapted to many different
ecological conditions and is one of the species becoming
increasingly dominant in forests due to its quick respond to
increases in light to the forest floor and rapid initial growth
rate [8]. Its increasingly important roles in forestry and wood
products is making studying Liriodendron of great interest.
Our study provides a first look at the genetic diversity and
allele richness among selections of this unique native species,
Table 5. Nei's (1978) unbiased identity (above diagonal) and
distance (below diagonal).

Clemson

Clemson Knoxville NA L.
chinense
-0.6792
0.6234

Hybrids in NA L.
Clemson chinense
0.6051
0.4051

Knoxville

0.3869

Pop ID

NA L.
0.4725
tulipifera
Hybrids in 0.5025
clemson
NA L.
0.9037
chinense

--

0.4648

0.4484

0.3495

0.7662

--

0.3608

0.3097

0.8021

1.0195

--

0.4138

1.0513

1.1721

0.8823

--

Supplementary Table S2
Supplementary figure S1
Supplementary figure S2
Supplementary figure S3

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions
Authors’ contributions

XZ AC ZT MS SES JEC HL

Research concept and design

--

--

--

--

--

✓

✓

Collection and/or assembly of data

✓

✓

✓

✓

✓

--

--

Data analysis and interpretation

✓

--

--

--

--

--

✓

Writing the article

✓

--

--

--

--

--

✓

Critical revision of the article

--

--

--

✓

✓

--

✓

Final approval of article

✓

✓

✓

✓

✓

✓

✓

Statistical analysis

✓

--

--

--

--

--

✓

Acknowledgement

The authors thank Nick Wheeler for his guidance and thorough
review of the manuscript. The funding for the study came from
the NSF Plant Genome Research program (NSF 1025974) and
National Institute of Food and Agriculture, USDA SC-1700449 with
a Clemson University Experiment Station technical contribution
number of 6215).

Publication history

Editor: Shouan Zhang, University of Florida, USA.
Received: 21-Jan-2015 Revised: 24-Feb-2015
Accepted: 17-Mar-2015 Published: 27-Mar-2015

References
1. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W and
Lipman DJ. Gapped BLAST and PSI-BLAST: a new generation of protein
database search programs. Nucleic Acids Res. 1997; 25:3389-402. |
Article | PubMed Abstract | PubMed Full Text
2. Bae K and Byun J. Screening of leaves of higher plants for antibacterial
action. Kor J Pharmacogn. 1987; 8:1.
3. Beck DE. Liriodendron tulipifera L. Yellow-poplar. In Silvics of North

7

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf
America. Harwoods. Edited by R.M. Burns and B.H. Honkala. US Dep
Agric Agric Handb 654. 1990; 2:406-416. | Article
4. Beckage B and Clark JS. Seedling survival and growth of three forest tree
328 species: The role of spatial heterogeneity. Ecology. 2003; 84:18491861. | Pdf
5. Bonner FT and Russell TE. Liriodendron tulipifera L. Yellow-poplar. In
Schopmeyer, CS (Tech. Coord.). Seeds of woody plants of the United
States. USDA For Serv Agric Handb. 1974; 450:508-511.
6. Cech F, Brown J and Weingartner D. Wind damage to a yellow-poplar
seed orchard. Tree Planters’ Notes. 1976; 27:3-4.
7. Dereeper A, Guignon V, Blanc G, Audic S, Buffet S, Chevenet F, Dufayard
JF, Guindon S, Lefort V, Lescot M, Claverie JM and Gascuel O. Phylogeny.
fr: robust phylogenetic analysis for the non-specialist. Nucleic Acids Res.
2008; 36:W465-9. | Article | PubMed Abstract | PubMed Full Text
8. Dyer JM. Using witness trees to assess forest change in southeastern
Ohio. Can J Bot. 2001; 31:1708-1718. | Article
9. Ellis JR and Burke JM. EST-SSRs as a resource for population genetic
analyses. Heredity (Edinb). 2007; 99:125-32. | Article | PubMed
10. Fageria MS and Rajora OP. Effects of silvicultural practices on genetic
diversity and population structure of white spruce in Saskatchewan.
Tree Gene Genome. 2014; 10:287-296. | Article
11. Forest Service. Forest Inventory and Analysis National Program, The
United State Department of Agriculture. FIADB 5.1.6 ed. 2014. | Website
12. Han Y, Chagne D, Gasic K, Rikkerink EH, Beever JE, Gardiner SE and
Korban SS. BAC-end sequence-based SNPs and Bin mapping for rapid
integration of physical and genetic maps in apple. Genomics. 2009;
93:282-8. | Article | PubMed
13. Hernandez R, Davalos JF, Sonti SS, Kim Y and Moody RC. Strength and
stiffness of reinforced yellow-poplar glued laminated beams. Res. Pap.
FPL-RP-U.S. Department of Agriculture, Forest Service, Forest Products
Laboratory, Madison, WI. 1997. | Pdf
14. Huang MR and Chen DM. An isozyme analysis of tulip-tree hybrids
(Liriodendron Chinese X L. tulipifera). J Nanjing For Univ. 1979; 1:156158. | Article
15. Jin H, Do J, Moon D, Noh EW, Kim W and Kwon M. EST analysis of functional genes associated with cell wall biosynthesis and modification in
the secondary xylem of the yellow poplar (Liriodendron tulipifera) stem
during early stage of tension wood formation. Planta. 2011; 234:959-77.
| Article | PubMed
16. Kimura M and Crow JF. The measurement of effective population number. Evolution. 1964; 17:279-288. | Article
17. Kobayashi N, Horikishi T, Katsuyama H, Handa T and Takayanagi K. A simple and efficient DNA extraction method for plants, especially woody
plants. Plant Tissue Culture Biotechnol. 1998; 4:76-80. | Article
18. Kovach KE. Assessment of genetic variation of Acer rubrum L. and
Liriodendron tulipifera L. populations in unmanaged forests of the
Southeast United States. 2009. | Article
19. Levene H. On a matching problem arising in genetics. Ann Math Stat.
1949; 20:91-94. | Article
20. Lewontin RC. The apportionment of human diversity. Evol Biol. 1972;
6:381-398. | Pdf
21. Li Z and Wang Z. RAPD markers used for the hybrid identification and
parents choice in Liriodendron. Sci Silv Sinic. 2002; 38:169-174. | Article
22. Liang H, Ayyampalayam S, Wickett N, Barakat A, Xu Y, Landherr L, Ralph
P, Xu T, Schlarbaum SE, Leebens-Mack JH and dePamphilis CW. Generation of a large-scale genomic resource for functional and comparative
genomics in Liriodendron. Tree Gen Genom. 2011; 7:941-954. | Article
23. Liang H, Carlson JE, Leebens-Mack JH, Wall PK, Mueller LA, Buzgo M,
Landherr LL, Hu Y, DiLoreto DS, Ilut DC, Field D, Tanksley SD, Ma H and dePamphilis CW. A 378 n EST Database for Liriodendron tulipifera L. floral
buds: the first EST resource for functional and comparative genomics in
Liriodendron. Tree Genet Genomes. 2008; 4:419-433. | Article
24. Liang H, Feng E, Tomkins J, Arumuganathan K, Zhao S, Luo M, Kudrna D,
Wing R, Banks J, dePamphilis C, Mandoli D, Schlarbaum S and Carlson JE.
Development of a BAC library resource for yellow poplar (Liriodendron

doi: 10.7243/2050-2389-4-1
tulipifera) and the identification of genomic regions associated with
flower development and lignin biosynthesis. Tree Genet Genomes. 2007;
3:215-225. | Article
25. Marshall TC, Slate J, Kruuk LE and Pemberton JM. Statistical confidence
for likelihood-based paternity inference in natural populations. Mol
Ecol. 1998; 7:639-55. | Article | PubMed
26. Merkle SA, Hoey MT, Watson-Pauley BA and Schlarbaum SE. Propagation
of Liriodendron hybrids via somatic embryogenesis. Plant Cell Tissue
Organ Cult. 1993; 34:191–198. | Article
27. Moon MK, Oh HM, Kwon BM, Baek NI, Kim SH, Kim JS and Kim DK. Farnesyl protein transferase and tumor cell growth inhibitory activities of
lipiferolide isolated from Liriodendron tulipifera. Arch Pharm Res. 2007;
30:299-302. | Article | PubMed
28. Nei M. Analysis of gene diversity in subdivided populations. Proc Natl
Acad Sci U S A. 1973; 70:3321-3. | PubMed Abstract | PubMed Full Text
29. Nei M. Estimation of average heterozygosity and genetic distance from
a small number of individuals. Genetics. 1978; 89:583-90. | Article |
PubMed Abstract | PubMed Full Text
30. Nurtjahjaningsih ILG, Saito Y, Tsuda Y and Ide Y. Genetic diversity of
parental and offspring populations in a Pinus merkusii seedling seed
orchard detected by microsatellite markers. Bulletin of the Tokyo University Forest, the Tokyo University Forests. 2007; 118:1-14. | Pdf
31. Parks CR, Miller NG, Wendel JF and Mc-Dougal KM. Genetic divergence
within the genus Liriodendron (Magnoliaceae). Ann Miss Bot Gard.
1983; 70:658-666. | Article
32. Parks CR and Wendel JF. Molecular divergence between Asian and North
American species of Liriodendron (Magnoliaceae) with implications for
interpretation of fossil floras. Am J Bot. 1990; 77:1243-1256. | Article
33. Raymond M and Rousset F. GENEPOP (version 1.2): population genetics
software for exact tests and ecumenisms. J Heredity. 1995; 86:248-249.
| Article
34. Santamour FS. Interspecific hybrids in Liriodendron and their chemical
verification. Fort Sci. 1972; 18:233-236. | Article
35. Sewell MM, Parks CR and Chase MW. Intraspecific chloroplast DNA
variation and biogeography of North American Liriodendron L. (Magnoliaceae). Evolution. 1996; 50:1147-1154. | Article
36. Thor E. Tree Breeding at the University of Tennessee, 1959-1976. Univ
Tenn Ag Exp Stn Bull, 1976; 48. | Article
37. Turmel M, Otis C and Lemieux C. The chloroplast genome sequence of
Chara vulgaris sheds new light into the closest green algal relatives of
land plants. Mol Biol Evol. 2006; 23:1324-38. | Article | PubMed
38. van Oosterhout C, Hutchinson WF, Wills DPM and Shipley P. MICROCHECKER: software for identifying and correcting genotyping errors in
microsatellite data. Mol Ecol Notes. 2004; 4:535-538. | Article
39. Wang Z. Utilization and species hybridization in Liriodendron. Chinese
Forestry Press, Beijing. 2005.
40. Wicke S and Quandt D. Universal primers for the amplification of the
plastid trnK/matK region in land plants. Anales Jard Bot Madrid. 2009;
66:285-288. | Article
41. Williams RS and Feist WC. Durability of yellow-poplar and sweetgum
and service life of finishes after long-term exposure. Forest Products J.
2004; 54:96-101. | Article
42. Xu M, Sun Y and Li H. EST-SSRs development and paternity analysis for
Liriodendron spp. New Forests. 2010; 40:361-382. | Article
43. Yang AH, Zhang JJ, Tian H and Yao XH. Characterization of 39 novel ESTSSR markers for Liriodendron tulipifera and cross-species amplification
in L. chinense (Magnoliaceae). Am J Bot. 2012; 99:e460-4. | Article |
PubMed
44. Yu JK, La Rota M, Kantety RV and Sorrells ME. EST derived SSR markers
for comparative mapping in wheat and rice. Mol Genet Genomics. 2004;
271:742-51. | Article | PubMed
45. Zane L, Bargelloni L and Patarnello T. Strategies for microsatellite isolation: a review. Mol Ecol. 2002; 11:1-16. | Article | PubMed
46. Zhang LJ, Oswald BP and Green TH. Relationships between overstory
species and community classification of the Sipsey Wilderness, Ala-

8

Zhang et al. Journal of Plant Science & Molecular Breeding 2015,
http://www.hoajonline.com/journals/pdf/2050-2389-4-1.pdf

doi: 10.7243/2050-2389-4-1

bama. For Ecol Manage. 1999; 114:377-383. | Article
47. Zhang H, Li H, Xu M and Feng Y. Identification of Liriodendron tulipifera,
Liriodendron chinense and hybrid Liriodendron using species-specific
SSR markers. Sci Silv Sinic. 2012; 46:36-39.

Citation:
Zhang X, Carlson A, Tian Z, Staton M, Schlarbaum SE,
Carlson JE and Liang H. Genetic characterization of
Liriodendron seed orchards with EST-SSR markers.
J Plant Sci Mol Breed. 2015; 4:1.
http://dx.doi.org/10.7243/2050-2389-4-1

9



Source Exif Data:
File Type                       : PDF
File Type Extension             : pdf
MIME Type                       : application/pdf
PDF Version                     : 1.4
Linearized                      : Yes
XMP Toolkit                     : Adobe XMP Core 5.3-c011 66.145661, 2012/02/06-14:56:27
Format                          : application/pdf
Rights                          : © 2012 Schmidt et al ; licensee Herbert Publications Ltd. This is an open access article distributed under the terms of  Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0),This permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Description                     : .
Title                           : Genetic characterization of Liriodendron seed orchards with EST-SSR markers
Creator                         : Haiying Liang
Subject                         : Genetic diversity, seed orchard, SSR markers, species
Create Date                     : 2015:03:28 09:55:24-07:00
Metadata Date                   : 2015:03:28 09:55:26-07:00
Modify Date                     : 2015:03:28 09:55:26-07:00
Rating                          : 5
Creator Tool                    : Adobe InDesign CS6 (Windows)
Instance ID                     : uuid:c95a2fb0-ba08-46fe-94ed-b761d4a14ba1
Original Document ID            : xmp.did:FD1BDF072C56E11188ADC19A7FDE6E7F
Document ID                     : xmp.id:BA7BEC1B6BD5E4119C60880B6C92D1EF
Rendition Class                 : proof:pdf
Derived From Instance ID        : xmp.iid:A5A1727C6AD5E4119C60880B6C92D1EF
Derived From Document ID        : xmp.did:E5FA261890C4E4119510948D51AB5E88
Derived From Original Document ID: xmp.did:FD1BDF072C56E11188ADC19A7FDE6E7F
Derived From Rendition Class    : default
History Action                  : converted
History Parameters              : from application/x-indesign to application/pdf
History Software Agent          : Adobe InDesign CS6 (Windows)
History Changed                 : /
History When                    : 2015:03:28 09:55:24-07:00
Marked                          : True
Web Statement                   : http://creativecommons.org/licenses/by-nc-nd/3.0/
Authors Position                : Journal of Plant Science & Molecular Breeding 2015, http://www.hoajonline.com/journals/pdf
Caption Writer                  : Research; Open access
Producer                        : Adobe PDF Library 10.0.1
Trapped                         : False
Page Count                      : 9
Author                          : Haiying Liang
EXIF Metadata provided by EXIF.tools

Navigation menu