ArticlePDF Available

Characterization and Application of EST-SSR Markers Developed From the Transcriptome of Amentotaxus argotaenia (Taxaceae), a Relict Vulnerable Conifer

Frontiers
Frontiers in Genetics
Authors:

Abstract and Figures

Amentotaxus argotaenia (Taxaceae) is a vulnerable coniferous species with preference for shade and moist environment. Accurate estimation of genetic variation is crucial for its conservation, especially in the context of global warming. In this study, we acquired a transcriptome from A. argotaenia leaves using Illumina sequencing and de novo assembled 62,896 unigenes, of which 5510 EST-SSRs were detected. Twenty-two polymorphic EST-SSRs were successfully developed and further used to investigate genetic variation, linkage disequilibrium, and bottleneck signatures of A. argotaenia. The results showed that A. argotaenia had moderate genetic variation and high genetic differentiation, which may provide raw material to protect against climatic changes and accelerate local adaptation, respectively. No bottlenecks were found to occur in A. argotaenia. Our study not only showed that these EST markers are very effective in population genetic analysis but also lay a solid foundation for further investigating adaptive evolution and conservation strategies of A. argotaenia.
Content may be subject to copyright.
1
Edited by:
Jacob A. Tennessen,
Harvard University,
United States
Reviewed by:
Jian-Feng Mao,
Beijing Forestry University,
China
Linkai Huang,
Sichuan Agricultural University,
China
*Correspondence:
Ting Wang
tingwang@scau.edu.cn
Yingjuan Su
suyj@mail.sysu.edu.cn
These authors have contributed
equally to this work
Specialty section:
This article was submitted to
Evolutionary and
PopulationGenetics,
a section of the journal
Frontiers in Genetics
Received: 31 March 2019
Accepted: 24 September 2019
Published: 18 October 2019
Citation:
RuanX, WangZ, WangT and
SuY (2019) Characterization and
Application of EST-SSR Markers
Developed From the Transcriptome of
Amentotaxus argotaenia (Taxaceae),
a Relict VulnerableConifer.
Front. Genet. 10:1014.
doi: 10.3389/fgene.2019.01014
Characterization and Application of
EST-SSR Markers Developed From
the Transcriptome of Amentotaxus
argotaenia (Taxaceae), a Relict
Vulnerable Conifer
Xiaoxian Ruan 1†, Zhen Wang 1†, Ting Wang 2* and Yingjuan Su 1,3*
1 School of Life Sciences, Sun Yat-sen University, Guangzhou, China, 2 College of Life Sciences, South China Agricultural
University, Guangzhou, China, 3 Research Institute of Sun Yat-sen University, Shenzhen, China
Amentotaxus argotaenia (Taxaceae) is a vulnerable coniferous species with preference
for shade and moist environment. Accurate estimation of genetic variation is crucial for
its conservation, especially in the context of global warming. In this study, we acquired a
transcriptome from A. argotaenia leaves using Illumina sequencing and de novo assembled
62,896 unigenes, of which 5510 EST-SSRs were detected. Twenty-two polymorphic
EST-SSRs were successfully developed and further used to investigate genetic variation,
linkage disequilibrium, and bottleneck signatures of A. argotaenia. The results showed
that A. argotaenia had moderate genetic variation and high genetic differentiation, which
may provide raw material to protect against climatic changes and accelerate local
adaptation, respectively. No bottlenecks were found to occur in A. argotaenia. Our study
not only showed that these EST markers are very effective in population genetic analysis
but also lay a solid foundation for further investigating adaptive evolution and conservation
strategies of A. argotaenia.
Keywords: Amentotaxus argotaenia, transcriptome, EST-SSR, genetic variation, population structure
INTRODUCTION
e genus Amentotaxus Pilger (Taxaceae) has six species, and ve of them have been listed as
endangered, vulnerable, or near threatened (Fu et al., 1999a; Farjon and Filer, 2013; Hilton-Taylor
et al., 2013). Among all Amentotaxus species, Amentotaxus argotaenia (Hance) Pilger has the widest
distribution but with small isolated populations occurring in southern and central China, northern
Vietnam, and Laos (Farjon and Filer, 2013). Its preferred environments are limestone mountains,
forests, ravines, and shady and damp stream banks, at altitudes of 300–1100 m (Fu et al., 1999a;
Lin et al., 2007). e natural regeneration of the plant is infrequent due to slow growth rate and
poorly dispersed seeds. Moreover, forest clearing and habitat loss have also been severely reducing
its population size. A. argotaenia is listed as vulnerable in China and as near threatened in the
International Union for Conservation of Nature Red List of reatened Species (Hilton-Taylor et al.,
2013). Since knowledge of population genetics is essential for the conservation and sustainable use
of wild resources (Wayne and Morin, 2004; Hoban et al., 2013), we aim to examine the population
genetic variation of A. argotaenia using novel molecular markers.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
ORIGINAL RESEARCH
doi: 10.3389/fgene.2019.01014
published: 18 October 2019
Transcriptome of Amentotaxus argotaeniaRuan et al.
2
In comparison to genomic SSRs, expressed sequence tag
(EST)-SSRs represent functional markers that may linked with
functional genes inducing phenotypic eects (Ranade et al.,
2014; Zhou et al., 2018a). ey hence provide opportunities to
examine functional diversity in relation to adaptive variation.
Moreover, EST-SSRs are reported to be more reliable because
they present lower frequencies of null alleles than do genomic
SSRs (Yu and Li, 2008). With the advances in next-generation
sequencing technology, EST-SSRs are becoming amenable to be
identied by sequencing the transcriptomes (Zhou et al., 2018b).
Next-generation sequencing is faster and more cost-eective than
traditional approach, e.g., cDNA library construction method
(Huang et al., 2015). However, currently, there is a shortage of
EST-SSRs developed from A. argotaenia, although EST-SSRs
have been identied for its congeneric species, Amentotaxus
formosana (Li et al., 2016). Excavation and characterization
of EST-SSRs for A. argotaenia may contribute to enhance our
understanding of its population genetic diversity, structure, and
the genetic basis of adaptive divergence. In addition, it will also
provide resources to assess the association between transposable
elements and SSR distribution as well as their roles in genome
organization (La Rota et al., 2005; Wang et al., 2018a).
In this study, we constructed a leaf transcriptome of A.
argotaenia using the Illumina sequencing platform. Based
on the transcriptome sequencing data, we developed a set of
EST-SSR markers and examined their polymorphisms. We
then assessed the genetic variation of four populations of A.
argotaenia using the novel EST-SSR markers. is work provides
essential information for the conservation and management of A.
argotaenia in the future.
MATERIALS AND METHODS
Plant Materials and DNA Extraction
A total of 56 A. argotaenia individuals were sampled from four
of its natural populations Jiuqushui (JQS; n = 15), Chuanping
(CP; n = 13), Qiniangshan (QNS; n = 12), and Wugongshan
(WGS; n = 16) located in China (Supplementary Table 1).
Fresh leaves were collected and desiccated in sealed plastic bags
with silica gel. Genomic DNA was isolated using the modied
cetyltrimethylammonium bromide method (Su et al., 2005). DNA
quality was evaluated using gel electrophoresis on 0.8% agarose gel.
RNA Extraction, cDNA Library
Construction, and Transcriptome
Sequencing
Fresh young leaves of one A. argotaenia individual from
population CP, which is planted in a greenhouse at Sun Yat-sen
University, were used to extract total RNAs using the method
described by Fu et al. (2004). RNA integrity was evaluated on
agarose gels followed by quantication on an Agilent 2100
Bioanalyzer (Agilent Technologies, Santa Clara, California,
USA). e mRNAs were isolated from the total RNAs by using
a Dynabeads mRNA DIRECT Kit (Invitrogen Life Technologies,
Carlsbad, California, USA) and randomly fragmented. e
fragmented mRNAs were converted into double-stranded cDNA
by using random primers and reverse transcriptase. Aer end-
repairing and tailing A, the cDNA fragments were ligated to
Illumina paired-end adapters. e cDNA library was sequenced
on an Illumina Hiseq2500 platform (Illumina, San Diego,
California, USA) with insertion size of 400–500 bp.
Transcriptome Assembly, Functional
Annotation, and Classification
We obtained a total of 25,257,542 paired-end reads from A.
argotaenia. e reads were ltered by removing primer or adaptor
sequences, and reads that contain unknown (“N”) or poor-
quality bases (the mean quality per base < 15 with a 4-base wide
sliding window) using the Trimmomatic soware version 0.32
(Bolger et al., 2014). e resulting clean data were deposited in
Sequence Read Archive of the National Center for Biotechnology
Information (NCBI) (Bioproject no. PRJNA413732; Biosample
no. SAMN07764634; https://www.ncbi.nlm.nih.gov/bioproject/
PRJNA413732). e clean reads longer than 90 nt were de novo
assembled into contigs and transcripts using the TRINITY
soware (https://github.com/trinityrnaseq/trinityrnaseq/releases)
with default settings. e transcripts that cannot be prolonged at
either end were dened as unigenes.
All unigenes were searched against Nt (NCBI nucleotide
sequences), Nr (NCBI non-redundant database), and Swiss-
Prot (a manually annotated and reviewed protein sequence
database) through blast 2.2.30+ (p://p.ncbi.nlm.nih.gov/blast/
executables/blast+/2.2.30/) with a cut-o E-value of 10−5. Protein
domains of open reading frame within unigenes was identied
by using HMMER hmmscan (hmmer-3.1b2-linux-intel-x86_64)
and Pfam (the protein families database). Blast2GO version 3.0
was used to perform Gene Ontology (GO) annotations dened by
molecular function, cellular component, and biological process
ontologies (http://www.blast2go.com/b2ghome). We further used
the Kyoto Encyclopedia of Genes and Genomes (KEGG) database
to performed pathways annotation and euKaryotic Ortholog
Groups (KOG) database to predict possible functions.
Development of EST-SSRs
EST-SSRs were searched in the assembled unigenes using
MIcroSAtellite (http://pgrc.ipk-gatersleben.de/misa/misa.html).
e SSRs were assumed to contain mono-, di-, tri-, tetra-, penta-,
and hexa-nucleotides with minimum repeat numbers of 10, 6, 5, 5,
5, and 5, respectively. Primer premier 5.0 (Clarke and Gorley, 2001)
was used to design primers. Only those SSRs containing two to six
repeat motifs were considered. Major parameters for primer design
included the following: (1) primer length ranging from 16 to 27
bp, (2) GC content of 30%–70%, (3) melting temperature between
50 and 63°C, and (4) PCR product size ranging from 100 to 400
bp. PCR reactions were performed in 20 μl mixture containing 2
μl 10 × PCR buer (Mg2+), 0.4 μl 10 mM dNTPs, 0.5 μl 10 mM
each of primers, 1.25 U Taq polymerase, and 20 ng DNA template.
e PCR protocol was as follows: 94°C for 5 min, followed by 30
cycles of 40 s at 94°C, 40 s at optimal annealing temperature and
30 s at 72°C, and a nal elongation at 72°C for 10 min. Amplified
products were screened on a 6.0% denaturing polyacrylamide gel,
and fragment size was determined with 50-bp marker.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
3
Evaluation of Polymorphic EST-SSRs and
Population Genetic Analysis
GenAlex software (Peakall and Smouse, 2006) was used
to calculate the EST-SSR genetic parameters, including
the number of observed alleles (Na), the effective number
of alleles (Ne), observed heterozygosities (Ho), expected
heterozygosities (He), and probability of the deviation from
the Hardy–Weinberg equilibrium. The polymorphism
information content (PIC) value and null alleles were
evaluated using CERVUS 3.0 (Kalinowski et al., 2007) and
Micro-Checker (Van Oosterhout et al., 2004), respectively.
Linkage disequilibrium across loci was determined using
TASSEL version 3.0 (Bradbury et al., 2007) with squared
correlation coefficient (r2 > 0.3) and the threshold of p values
(< 0.001) based on Fisher’s exact test.
Arlequin version 3.5 (Excoer and Lischer, 2010) was used to
perform a Mantel test with 10,000 permutations to examine the
pattern of isolation by distance. Using the same soware, analysis
of molecular variance was conducted to determine the amount of
genetic variation at dierent levels.
Using EST-SSR data, we applied a Bayesian model-based
clustering algorithm implemented in STRUCTURE 2.2 to infer
population structure. We used the admixture model, setting
the parameters as follows: burn-in periods = 10,000, MCMC
replicates = 10,000, K = 1 to 8, and iterations = 10. e optimum
number of clusters (K) was determined by calculating ΔK
(Evanno et al., 2005).
We conducted the Wilcoxons sign-rank test and the mode-
shi test to detect signatures of genetic bottleneck by running
BOTTLENECK version 1.2.02 (Piry et al., 1999). e two-phase
mutation model was selected because it is more suitable for
microsatellite data than the other two models (Piry et al., 1999;
Zhang and Zhou, 2013). We performed 1000 simulations under
the two-phase mutation model with 70% single-step mutations
and 30% multi-step mutations.
RESULTS
Transcriptome Sequencing and
DeNovoAssembly
Approximately 23.5 million clean reads were obtained
from the transcriptome of A. argotaenia with the length of
90–125 bp and GC content of 47%. The percentages of Q20
(base sequencing error probability < 1%) and Q30 (base
sequencing error probability < 0.1%) bases were 100% and
97%, respectively. These clean reads were assembled into
80674 transcripts by using Trinity with an average length of
756 bp and an N50 of 1018 bp. The total length of transcripts
reached 60,999,479 bp. After further assembly, a total of
62,896 unigenes were identified with an average length of
721 bp, a minimal length of 301 bp, and an N50 value of 947
bp. The sum of the length of the unigenes was 45,357,136 bp
(Table 1). The length of 37.473% (23569) of the unigenes
ranged from 301 to 400 bp, 61.956% (38,968) varied from
401 to 3000 bp, while 0.571% (359) was longer than 3000 bp
(SupplementaryFigure 1).
Functional Annotation and Categorization
We conducted the annotation of 62,896 unigenes in the seven
public databases (Nr, Nt, KOG, Swiss-Prot, Pfam, KEGG, and
GO), of which 36,671 were successfully annotated (Table 1). Of
them, 13,140 (20.89%) were annotated in Nt, 32,183 (51.17%)
in Nr, 26,309 (41.83%) in Swiss-Prot, 21,595 (34.33%) in Pfam,
31,283 (49.74%) in GO, 32,953 (52.39%) in KOG, and 15,072
(23.96%) in KEGG. A total of 490 unigene were identied in all
seven databases (Tabl e 1).
First, 31,283 unigenes were classied into three main GO
categories: biological process, cellular component, and molecular
function, including 44 functional groups. In the biological process
category, there were 7313 unigenes assigned to “cellular process,
5973 to “single-organism process,” 5808 to “metabolic process,
and 1 to “biological regulation.” In the cellular component
category, “cell part” and “organelle” component-related functions
were predominant, with 4835 unigenes assigned to the former and
2060 to the latter. In the molecular function category, “binding”
and “catalytic activity” were the most enriched, comprising 2520
and 2358 unigenes, respectively (Figure 1).
Second, 32,953 unigenes were assigned to 25 KOG classications.
Among them, the largest group was “general function prediction only”
(4286), followed by “function unknown” (3218), “signal transduction
mechanisms” (2521), and “posttranslational modication, protein
turnover, chaperones” (2442) (Supplementary Figure 2).
ird, a total of 15,072 unigenes were mapped to 38 KEGG
pathways corresponding to six categories (Supplementary Figure
3). Most of them involved in “translation” (3007 unigenes; 19.95%),
carbohydrate metabolism” (2815; 18.68%), and “folding” (2425;
16.09%) pathways. e 38 KEGG pathways were related to such
categories as metabolism, cellular processes, genetic information
processing, environmental information processing, and others.
TABLE 1 | Sequencing, assembly, and annotation results of Amentotaxus
argotaenia transcriptome.
Total raw reads 25,257,542
Total clean reads 23,553,846
% Q20 100
% Q30 97
% GC 47
Number of transcripts 80,674
Average length of transcripts (bp) 756
N50 of transcripts (bp) 1018
Number of unigenes 62,896
Minimum length of unigenes (bp) 301
Average length of unigenes (bp) 721
N50 of unigenes (bp) 947
Number of unigenes annotated at least one databases 36,671
Number of unigenes with Nt annotations 13,140
Number of unigenes with Nr annotations 32,183
Number of unigenes with KOG annotations 32,953
Number of unigenes assigned to GO terms 31,283
Number of unigenes with Swiss-Prot annotations 26,309
Number of unigenes with Pfam annotations 21,595
Number of unigenes with KEGG pathways 15,072
Number of unigenes annotated to seven public databases 490
Nt, NCBI nucleotide sequences; Nr, NCBI non-redundant database; KOG,
euKaryotic Ortholog Groups; GO, Gene Ontology; KEGG, Kyoto Encyclopedia of
Genes and Genomes.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
4
Characterization of EST-SSRs
We identied 5510 EST-SSRs from 4830 SSR-containing unigene
sequences, of which 362 were compound SSRs. With the
exception of mononucleotide repeats (64.50%), trinucleotide
repeats were the most common type with a frequency of 22.25%,
followed by di- (9.18%), tetra- (1.82%), penta- (1.00%), and
hexanucleotide (1.25%) repeats (Table 2). EST-SSRs with 10
tandem repeats (1719, 31.19%) were the most common, followed
by >11 tandem repeats (1134; 20.58%) and 9 tandem repeats
(42; 0.76%) (Table 2). Motif A/T (98.11%) was dominant in
TABLE 2 | The distribution of repeat unit number and motif length of EST-SSRs.
SSR motif length Repeat unit number
5 6 7 8 9 10 11 > 11 Total Percentage (%)
Mono 1682 787 1085 3554 64.50
Di 255 89 60 32 20 16 34 506 9.18
Tri 747 266 112 61 8 15 5 12 1226 22.25
Tetra 65 27 1 4 1 2 100 1.82
Penta 42 3 3 3 2 2 55 1.00
Hexa 34 18 5 8 1 2 1 69 1.25
Total 888 569 210 136 42 1719 812 1134 5510
Percentage (%) 16.12 10.33 3.81 2.47 0.76 31.19 14.74 20.58
Mono, mono-nucleotide; Di, di-nucleotide; Tri, tri-nucleotide; Tetra, tetra-nucleotide; Penta, penta-nucleotide; Hexa, hexa-nucleotide.
FIGURE 1 | Functional classification of Gene Ontology for Amentotaxus argotaenia unigenes.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
5
mononucleotide repeats, whereas AT/TA (50.79%) was the most
abundant in dinucleotide repeats, followed by AG/CT (35.18%)
and AC/GT (14.03%). Trinucleotide motifs AAG/CTT, AGG/
CCT, AGC/CTG, AAT/ATT, and ACT/AGT had frequencies of
21.21%, 17.94%, 15.42%, 14.68%, and 0.33%, respectively.
We also examined the linkage of the 22 EST-SSR loci. Of 231
combinations, four pairs of loci (SHS-1306 vs SHS-26840, SHS-
1306 vs SHS-38297, SHS-28717 vs SHS-32587, and SHS-34629 vs
SHS-38297) were found to be linked with r2 > 0.3 and p < 0.001.
Polymorphic EST-SSRs Identification and
Estimation of Genetic Diversity
We randomly selected 60 EST-SSR primers to evaluate their
application and the polymorphism across 12 A. argotaenia
individuals from four populations. Twenty-two of the
microsatellite loci exhibited allelic polymorphism, whereas 16
were identied as monomorphic (Tabl e 3). We then used the 22
polymorphic EST-SSR markers to perform population genetic
analysis (Supplementary Figure 5).
e number of observed alleles and the eective number of
alleles varied from 1 to 7 and from 1 to 4.694 per locus, respectively.
e observed heterozygosity ranged from 0 to 1.000 (average =
0.250), while the expected heterozygosity ranged from 0.000 to
0.787 (average = 0.390). e mean value of PIC was 0.455, with
the minimum of 0.084 and the maximum of 0.707. Fourteen loci
were identied as null allele. Six, 9, 14, and 16 EST-SSRs showed
signicant deviations from the Hardy–Weinberg equilibrium in
populations JQS, CP, QNS, and WGS, respectively (Table 4).
Population Genetic Structure
andDifferentiation
Genetic dierentiation (Fst) based on EST-SSRs was 0.28198.
Analysis of molecular variance revealed that 71.80% of the genetic
variation occurred within populations, while 10.37% and 17.83%
were attributed to among populations within groups and among
groups, respectively (Supplementary Table 2). In addition, the
result of Mantel test showed that there was no signicant correlation
between genetic and geographical distances (p = 0.3306).
ΔK demonstrated that the uppermost K equaled 3
(Supplementary Figure 4). Amentataxus argotaenia populations
were assigned to three groups. Group I contained populations
JQS and CP, group II contained population WGS, and group III
contained population QNS (Figure 2).
Bottleneck Signature
At species level, both Wilcoxons sign-rank test and mode-shi
analysis indicated that A. argotaenia had not experienced a
recent bottleneck (Supplementary Table 3).
DISCUSSION
Characterization of Transcriptome
In this study, we sequenced, assembled, and annotated the
transcriptome of A. argotaenia using the next-generation
sequencing approach. A total of 62,896 unigenes were de novo
assembled with the unigene mean and N50 length of 721 and 947
bp, respectively (Table 1). More than half of the unigenes can be
successfully annotated through seven databases (Nr, Nt, KOG,
Swiss-Prot, Pfam, KO, and GO), of which 490 were simultaneously
identied (Tab le 1 ). Most of the annotated unigenes were unique in
A. argotaenia compared to its closely related conifer Torreya grandis
(Zeng et al., 2018). Moreover, the annotated results of NR database
indicated that A. argotaenia exhibited only 27.5% unigene identity
to another conifer Picea sitchensis. ese unique unigenes may
represent the species-specic genetic signature of A. argotaenia
potentially underlying its speciation process or evolution (Frech
and Chen, 2011). Similar results have been found in the case of
Picea abies (Nystedt et al., 2013). In addition, 31,283, 32,953, and
15,072 unigenes of A. argotaenia were assigned into 44 functional
groups in GO, 25 classications in KOG, and 38 pathways in KEGG,
respectively. ese results indicate that the identied A. argotaenia
unigenes have wide-ranging functions and will be valuable for
analyzing the functional diversity of A. argotaenia.
Frequency and Distribution of EST-SSRs
e mean length of unigenes (721 bp) of A. argotaenia was
considerably longer than that of other conifers, including Pinus
pinaster (495 bp) (Canales et al., 2014), Platycladus orientalis (686
bp) (Hu et al., 2016), and P. a bie s (472 bp) (Chen et al., 2012).
Zalapa et al. (2012) pointed out that longer sequences will increase
the probability to properly design EST-SSR primers. In accordance
with this, we developed 5510 EST-SSRs from the transcriptome
of A. argotaenia, which was signicantly greater than those of A.
formosana (4955) (Li et al., 2016) and Pinus densiora (1953) (Liu et
al., 2015). e most common motif for dinucleotides in A. argotaenia
was found to be AT/TA. e same result was obtained in Picea
spp. (Rungis et al., 2004), Pinus dabeshanensis (Xiang et al., 2015),
and Pseudolarix amabilis (Geng et al., 2015). Ranade et al. (2014)
emphasized that AT/TA was oen ranked as the most abundant
dimer motif in gymnosperms (especially in the 3 untranslated
regions). A factor related to this phenomenon may be increased A +
T contents. In addition, the most common trinucleotide motif in A.
argotaenia was AAG/CTT, which is similar to that in Cryptomeria
japonica (Ueno et al., 2012), Pinus taeda (Wegrzyn et al., 2014), and
Pinus halepensis (Pinosio et al., 2014), but unlike in P. dabeshanensis
(AGC) (Xiang et al., 2015). It has been noted that AAG/CTT is the
target for methylation in plants (Law and Jacobsen, 2010).
Validation and Polymorphism of EST-SSRs
and Population Genetic Variation
irty-eight of the 60 EST-SSR primers designed for A. argotaenia
enabled to amplify expected products with a success rate of 63.33%,
which is relatively high in comparison to previous studies (Dantas
et al., 2015; Ueno et al., 2015). en 22 polymorphic EST-SSRs
were used to investigate standard genetic diversity and population
genetic structure of four A. argotaenia populations. Only four pairs
of them were tightly linked among each other. Moderate genetic
variation was observed based on the EST-SSRs (PIC = 0.455)
according to the evaluation criteria of polymorphism (moderate:
0.25 < PIC < 0.5) (Botstein et al., 1980). is judgment was also
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
6
TABLE 3 | Characterization of 38 EST-SSR primer pairs for A. argotaenia.
Locus Primer sequences (5–3)Repeat
motif
Allele
size(bp)
Ta(°C) GenBank
accession no.
Putative function
[Organism]
E-value
SHS-1306 F: ACCTCGGGTCCTGTTGAAR:
GGTTGTGGCGAATGCTG
(TAC)6(TAT)5249 54 MG209531 Unknown [Picea sitchensis] 3e-53
SHS-1845 F: TCTGAGATAAGGTGCTTGGGTGR:
ATTTGAGGGCTACAGCGGTT
(GTC)7126 55 MG209532 Unknown [P. sitchensis] 3e-48
SHS-2019 F: AGAGACCACCAACGACGAACR:
CAGCGGCAGCATACCATT
(TCC)7298 55 MG209533 hypothetical protein AMTR_
s00032p00067030 [Amborella
trichopoda]
3e-25
SHS-2589 F: CTAACCCTATCCCTAACCTCTTTCR:
GTTTCATTCCAGGCACTCTCA
(GAA)5159 57 MG209534 Unknown [P. sitchensis] 2e-121
SHS-5811 F: TAGATTTAGTTCCCAGCGGTGR:
GATTGATTTCGGCTCGTGTAT
(AG)8279 53 MG209535 Not found
SHS-6181 F: TTCTACTTCTGCTGCTGGTGTR:
GCATTGGTCTTCCTCCTTTAC
(TTC)7214 54 MG209536 hypothetical protein
AMTR_s00003p00222410 [A.
trichopoda]
0
SHS-17706 F: CTCTTTGGGAGAAGTATTAGCR:
TGGTCACTCGTGGACATTA
(AAGA)5134 53 MG209537 PREDICTED: uncharacterized
protein LOC103491482
isoform X2 [Cucumis melo]
5e-45
SHS-18170 F: GAGAGCCCACGGTCCTGTR:
AGTCCCATCATCCACCTATCA
(GAA)6300 53 MG209538 PREDICTED: zinc finger
protein 43-like [Pyrus x
bretschneideri]
1e-10
SHS-18213 F: AAAGTCGGGTGATTACAGAGCR:
TCCTTCGTGGAATGTTTATGA
(GA)7397 54 MG209539 Not found
SHS-18563 F: ACCTCCTACACCCCCTTCTR:
AACTCCACCATACGCATCTTA
(GAA)6236 54 MG209540 Unknown [P. sitchensis] 6e-97
SHS-19735 F: CCCAAAGAAAGGGCAAGAR:
CGGCGGATGGTAATGTG
(CGC)7278 53 MG209541 Unknown [P. sitchensis] 3e-64
SHS-20137 F: CTGTCAGGCATTTCTGGGTCTR:
CGATTTTCATTTTGTTTGGTCTG
(CTT)7257 58 MG209542 Not found
SHS-20198 F: CATTCTCACACCCTTGTATTGCTR:
CATCTTCACCATTTCTCTGTAGTCTT
(TA)8259 58 MG209543 transcription factor AP2 [Taxus
cuspidata]
8e-117
SHS-21264 F: CTCGTCCAAGAAGAACCATACR:
CATCATAAACCACTTAGCAAATAC
(GAG)6400 56 MG209544 PREDICTED: uncharacterized
protein LOC104240103
[Nicotiana sylvestris]
7e-41
SHS-21490 F: GAGGAAGAGGGTTTTGGTCATR:
AGTAGGCGTCTTTGGCGTT
(TAA)6190 58 MG209545 hypothetical protein PRUPE_
ppa010075mg [Prunus
persica]
1e-59
SHS-22515 F: CACATCCTCCGCCGACTR:
TTGCTGTTTTACCGAGAAGAAG
(TAC)6(TAT)5266 57 MG209546 Unknown [P. sitchensis] 2e-53
SHS-23191 F: ACCCAGTTGTGGTAGGAGCATR:
AAAGTGTGAAACATCCCAAAGC
(GAG)6161 57 MG209547 Not found
SHS-23195 F: TGACAACGAGAACGAAGAACATAACR:
GTCTGTAAGCCAACGCTGAGG
(AGA)6115 57 MG209548 hypothetical protein
SELMODRAFT_451322
[Selaginella moellendorffii]
2e-41
SHS-24187 F: CCTAATGGTGAATAACTTGTGCTCR:
GCGAGTTTCTTGCTAAATGCTT
(TCTT)5330 58 MG209549 Not found
SHS-24301 F: TACCTGACTGGACTGCTGAATR:
ATGTTAGAGGAATACGATAGGCT
(CCG)6377 57 MG209550 Unknown [P. sitchensis] 1e-32
SHS-26622 F: AGATACTCTTGTTTCAGGAGCATTR:
CAACCCAGGACATCACCATAG
(AAG)7228 57 MG209551 Not found
SHS-26840 F: GGGCGGAGGAGAATGGTCR:
TGGGCTGCTGAAATAGGAAAC
(GGA)6250 57 MG209552 PREDICTED: uncharacterized
protein LOC104602979,
partial [Nelumbo nucifera]
1e-65
SHS-28207 F: CAATCGGATAAGGTGTTTCTR:
CGAATAGTGGTAATCAAATAGG
(GA)16 379 52 MG209553 Unknown [P. sitchensis] 0
SHS-28326 F: GTATGGAAGGGAGGCGAAATR:
GCCGCTGTGGTTGTGAAG
(ATA)6143 57 MG209554 Unknown [P. sitchensis] 4e-29
SHS-28474 F: AATAAGAATAGGAGGGGTGAAGACR:
GAGACAGAGGATTTGTAACGGAG
(CGG)6218 57 MG209555 PREDICTED: mediator of RNA
polymerase II transcription
subunit 30-like isoform X1 [N.
nucifera]
1e-16
SHS-28717 F: CGTATCCCTGTTGATTCATTTTCR:
GGTTGTATCATTCAGTCCCATTG
(GAAA)6240 57 MG209556 Not found
(Continued)
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
7
TABLE 4 | Genetic diversity statistics for four A. argotaenia populations based on 22 polymorphic EST-SSR primers.
JQS (N = 15) CP (N = 13) QNS (N = 12) WGS (N = 16)
Locus NaNeHoHeNaNeHoHeNaNeHoHeNaNeHoHePIC
SHS-1306 2 1.923 0.000*0.480 2 1.166 0.000* 0.142 2 1.180 0.000* 0.153 2 1.600 0.000*0.375 0.395
SHS-2019 2 1.301 0.267 0.231 3 1.807 0.462 0.447 2 1.180 0.000* 0.153 2 1.133 0.000* 0.117 0.429
SHS-6181 1 1.000 0.000 0.000 1 1.000 0.000 0.000 2 1.087 0.083 0.080 3 2.327 0.000*0.570 0.225
SHS-17706 3 1.822 0.600 0.451 4 1.931 0.2310.482 3 2.880 0.500* 0.653 3 2.667 0.500* 0.625 0.605
SHS-18213 4 1.531 0.000*0.347 2 1.742 0.000*0.426 4 1.714 0.000*0.417 3 1.910 0.000*0.477 0.402
SHS-18563 1 1.000 0.000 0.000 1 1.000 0.000 0.000 2 1.800 0.000*0.444 3 2.844 0.000*0.648 0.371
SHS-19735 4 2.217 0.600 0.549 2 1.352 0.000*0.260 2 1.385 0.000*0.278 3 1.684 0.250* 0.406 0.393
SHS-20198 2 1.923 0.000*0.480 3 1.610 0.000*0.379 4 3.236 0.250*0.691 3 2.032 0.000*0.508 0.560
SHS-24187 4 2.284 0.467 0.562 3 2.048 0.692 0.512 3 2.667 0.667 0.625 4 2.090 0.375* 0.521 0.528
SHS-26622 6 4.091 0.800 0.756 7 4.694 1.000 0.787 4 2.743 1.000 0.635 4 2.498 1.000 0.600 0.707
SHS-26840 4 2.761 0.867 0.638 4 2.467 0.462 0.595 3 2.268 0.583 0.559 2 2.000 1.000* 0.500 0.514
SHS-28207 4 2.663 1.000 0.624 5 3.282 1.000 0.695 5 3.840 0.917* 0.740 3 1.684 0.500 0.406 0.701
SHS-28474 4 1.772 0.000*0.436 3 1.610 0.000*0.379 3 2.323 0.000*0.569 4 1.707 0.000*0.414 0.456
SHS-28717 2 1.471 0.000*0.320 2 1.742 0.000*0.426 3 2.323 0.000*0.569 2 1.280 0.000*0.219 0.581
SHS-31463 1 1.000 0.000 0.000 2 1.257 0.231 0.204 1 1.000 0.000 0.000 2 1.133 0.125 0.117 0.084
SHS-31908 1 1.000 0.000 0.000 3 1.857 0.000*0.462 2 1.800 0.000*0.444 3 1.293 0.000*0.227 0.297
SHS-32587 2 1.471 0.000*0.320 3 1.610 0.000*0.379 4 2.595 0.083*0.615 3 2.612 0.000*0.617 0.534
SHS-32686 2 1.800 0.267 0.444 2 1.352 0.308 0.260 2 1.180 0.000* 0.153 2 1.969 0.1250.492 0.399
SHS-32939 3 1.495 0.400 0.331 2 1.451 0.385 0.311 2 1.180 0.167 0.153 2 1.992 0.563 0.498 0.393
SHS-34629 2 1.301 0.267 0.231 2 1.649 0.538 0.393 2 1.385 0.000*0.278 2 1.600 0.000*0.375 0.473
SHS-35453 2 1.800 0.667 0.444 2 1.988 0.000*0.497 2 1.882 0.583 0.469 3 2.667 0.500* 0.625 0.579
SHS-38297 2 1.301 0.267 0.231 2 1.451 0.385 0.311 2 1.087 0.083 0.080 1 1.000 0.000 0.000 0.384
N, number of individuals of each populations; Na, number of alleles per locus; Ne, number of effective alleles per locus; Ho, observed heterozygosity; He, expected
heterozygosity; PIC, polymorphism information content; *significant deviation from the Hardy–Weinberg equilibrium (p < 0.001); significant possibility of the presence
ofnull alleles detected by MICRO-CHECKER.
TABLE 3 | Continued
Locus Primer sequences (5–3)Repeat
motif
Allele
size(bp)
Ta(°C) GenBank
accession no.
Putative function
[Organism]
E-value
SHS-30422 F: TTCCTTCTACTCCCTCTTCTATGTCR:
AACTGCTTACCTAAATGGTGCTG
(CTT)6163 57 MG209557 PREDICTED: probable
cellulose synthase A catalytic
subunit 5 [Phoenix dactylifera]
0
SHS-31463 F: ATGGATGGCAGGATTGGATR:
AACAAATAAGGAAGAAGGTGGTAGT
(TTG)6182 57 MG209558 Not found
SHS-31820 F: TTTGGTTCCATACCTGCTCCTR:
TTCGTGGTCACTCTTTTCCCT
(GAT)6149 57 MG209559 Unknown [P. sitchensis] 5e-172
SHS-31908 F: CCAGACTTGCCACATCAGCR:
AACCCACAACCCACCAGAG
(ATG)5(AGG)7395 57 MG209560 Unknown [P. sitchensis] 4e-11
SHS-32587 F: AAATGAGGAATAAGTAGGTGAAGTTR:
GCACATTAGGGTTCCTGATTAC
(AGA)6254 57 MG209561 Not found
SHS-32686 F: CAACCCGTCCCTTGCTTTAGR:
CCTCTGCGTCCTTGTTGTTATC
(GGCAG)5302 57 MG209562 Unknown [P. sitchensis] 6e-176
SHS-32939 F: TGGAAAAAACCACAGACGACTCR:
GCCCTCAAACACAAAAGCAG
(ATA)7143 56 MG209563 Unknown [P. sitchensis] 1e-94
SHS-34629 F: CTGGACAAAGAGAGCAACGGTR:
AATGGCGACACAAGTGAGAAGT
(CTC)7222 56 MG209564 Hypothetical protein
AMTR_s00032p00067030 [A.
trichopoda]
2e-19
SHS-35222 F: TGCTGCCTAAACACAATGTCTCTR:
CACAAGTCTTCCTTTTCCCTAATG
(TG)9121 56 MG209565 Not found
SHS-35453 F: GTTGAGCATTGATTTAGATGTTCGR:
TTTCCCCTCCTCTTTCTTTGAC
(TGA)8189 55 MG209566 PREDICTED: uncharacterized
protein LOC104596414
isoform X2 [N. nucifera]
2e-68
SHS-38297 F: TTACCAACGCCAAATGCTGR:
ACCCTACTCCCACTCCCTTCT
(GAGATG)7154 55 MG209567 hypothetical protein
AMTR_s00099p00142540 [A.
trichopoda]
1e-24
SHS-39519 F: TTGTGCCTCTTCAAGGAGTAGTR:
GAGAATCTTCCCTGTCGGTC
(CTC)6261 55 MG209568 ACC synthase-like [Picea
glauca]
0
Ta, annealing temperature; bold font, polymorphic loci.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
8
lent support by such genetic parameters as the average of observed
alleles, observed heterozygosity, expected heterozygosity, and the
percentage of polymorphic band. In contrast, previous researches
detected low levels of genetic diversity in A. argotaenia by using
ISSRs and genomic SSRs (Ge et al., 2005; Ge et al., 2015). ese
inconsistencies highlighted the importance to investigate genetic
variation of A. argotaenia using multiple markers.
Accurate estimate of genetic diversity is very useful for
conservation and management of genetic resources (Cardoso et al.,
1998; Wang et al., 2018b). Compared to other conifers, we observed
a moderate EST-SSR variation in A. argotaenia (Supplementary
Tabl e 4). Similar levels of the EST-SSR variation were found in its
closely related species A. formosana as well (Li et al., 2016). e
moderate level of functional diversity, together with the nding
that A. argotaenia did not experience a recent bottleneck, implies
that the species still has essential evolutionary potential to adapt to
the changing environment (Frankham, 2010).
We detected a marked genetic dierentiation among
A. argotaenia populations in comparison to other conifers
(Supplementary Table 4). Similar ndings were obtained by
using chloroplast intergenic spacer, mitochondrial intron, and
genomic microsatellite data (Ge et al., 2015). e dispersal
distance of pollen and seed in conifers is generally less than 2 and
20 km, respectively (Fu et al., 1999b; Cain et al., 2000; Sima, 2004;
Zhang et al., 2005; Lu, 2006). As for A. argotaenia, its pollen and
seed exchanges may be further hindered because of preferring
to grow under forest canopies (Ge et al., 2015). It is thus
reasonable to speculate that the genetic dierentiation pattern of
A. argotaenia is highly linked to restricted between-population
gene ow (genetic exchange via pollen and seed). Moreover, the
establishment of climate- or habitat-linked genotypes should
also be considered, since we used functional markers to perform
studies (Jump and Peñuelas, 2005; Ortego et al., 2012).
CONCLUSION
We generated the leaf transcriptome of A. argotaenia by using
Illumina sequencing technology. A total of 62,896 unigenes were
assembled, annotated, and classied. Based on the transcriptome
data, 5510 EST-SSRs were identied from 4830 SSR-containing
unigene sequences. Among them, 60 were randomly selected for
the development of potential functional markers. Consequently,
22 polymorphic EST-SSR markers were developed and used
to reveal a moderate level of functional diversity, along with
marked genetic structure and the lack of genetic bottleneck, in A.
argotaenia. is study has provided eective EST-SSR markers for
measuring the evolutionary potential of A. argotaenia in response
to environmental changes.
DATA AVAILABILITY STATEMENT
All Illumina clean data generated for this study was deposited
at the Sequence Read Archive (SRA) of the National Center for
Biotechnology Information (https://www.ncbi.nlm.nih.gov/sra/
SRX3296043[accn]). e Bioproject number and Biosample
number for clean data are PRJNA413732 and SAMN07764634,
respectively. irty-eight EST-SSR sequences generated for
this study were deposited at GenBank with accession numbers
MG209531-MG209568.
AUTHOR CONTRIBUTIONS
e author XR conducted the experiments and ZW completed
the data analysis. e two authors contributed equally to this
work. YS designed the experiments and wrote the manuscript,
and TW corrected the manuscript.
FUNDING
is work was supported by the National Natural Science
Foundation of China (31370364, 31570652, 31670200, 31770587,
and 31872670); the Natural Science Foundation of Guangdong
Province, China (2016A030313320 and 2017A030313122);
Science and Technology Planning Project of Guangdong
Province, China (2017A030303007); Project of Department
of Science and Technology of Shenzhen City, Guangdong,
China (JCYJ20160425165447211, JCYJ20170413155402977,
and JCYJ20170818155249053); and Science and Technology
Planning Project of Guangzhou City, China (201804010389).
ACKNOWLEDGMENTS
The authors thank Dr. Q. Fan of the School of Life Sciences,
Sun Yat-sen University, for assistance with the collection of
plant materials.
SUPPLEMENTARY MATERIAL
e Supplementary Material for this article can be found online at:
https://www.frontiersin.org/articles/10.3389/fgene.2019.01014/
full#supplementary-material
SUPPLEMENTARY FIGURE 1 | Length distribution of assembled unigenes
generated from A. argotaenia transcriptome.
FIGURE 2 | Population genetic structure analysis of A. argotaenia based on
EST-SSRs. JQS, Population Jiuqushui; CP, Population Chuanping; WGS,
Population Wugongshan; QNS, Population Qiniangshan.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
9
SUPPLEMENTARY FIGURE 2 | Functional classification of A. argotaenia
unigenes based on KOG annotation.
SUPPLEMENTARY FIGURE 3 | Functional classification of A. argotaenia
unigenes based on KEGG annotation.
SUPPLEMENTARY FIGURE 4 | The optimum number of clusters (K) estimated
with STRUCTURE analysis based on LnP(D) (A) and hoc statistic ΔK (B).
SUPPLEMENTARY FIGURE 5 | Amplification products generated using
EST-SSR primer pair, SHS-26622, were separated by electrophoresis in 6%
denaturing polyacrylamide gel. The expected allele size was 228 bp and the
annealing temperature was 57°C. Lanes 1-15, 16-28, 29-40, and 41-56 were
products of A. argotaenia individuals from populations JQS (Jiuqushui), CP
(Chuanping), QNS (Qiniangshan), and WGS (Wugongshan), respectively; Lane M:
50 bp ladder.
REFERENCES
Bolger, A. M., Lohse, M., and Usadel, B. (2014). Trimmomatic: a exible trimmer
for Illumina sequence data. Bioinformatics 30, 2114–2120. doi: 10.1093/
bioinformatics/btu170
Botstein, D., White, R. L., Skolnick, M., and Davis, R. W. (1980). Construction of a
genetic linkage map in man using restriction fragment length polymorphisms.
Am. J. Hum. Genet. 32, 314–331. doi: 10.1016/0165-1161(81)90274-0
Bradbury, P. J., Zhang, Z. W., Kroon, D. E., Casstevens, T. M., Ramdoss, Y., and
Buckler, E. S. (2007). TASSEL: soware for association mapping of complex
traits in diverse samples. Bioinformatics 23, 2633–2635. doi: 10.1093/
bioinformatics/btm308
Cain, M. L., Milligan, B. G., and Strand, A. E. (2000). Long-distance seed dispersal
in plant populations. Am. J. Bot. 87, 1217–1227. doi: 10.2307/2656714
Canales, J., Bautista, R., Label, P., Gomez-Maldonado, J., Lesur, I., Fernandez-
Pozo, N., et al. (2014). De novo assembly of maritime pine transcriptome:
implications for forest breeding and biotechnology. Plant Biotechnol. J. 12,
286–299. doi: 10.1111/pbi.12136
Cardoso, M. A., Provan, J., Powell, W., Ferreira, P. C. G., and De Oliveira, D.
E. (1998). High genetic dierentiation among remnant populations of the
endangered Caesalpinia echinata Lam. (Leguminosae–Caesalpinioideae). Mol.
Ecol. 7, 601–608. doi: 10.1046/j.1365-294x.1998.00363.x
Chen, J., Uebbing, S., Gyllenstrand, N., Lagercrantz, U., L ascoux, M., and Källman,
T. (2012). Sequencing of the needle transcriptome from Norway spruce
(Picea abies Karst L.) reveals lower substitution rates, but similar selective
constraints in gymnosperms and angiosperms. BMC Genomics 13, 589. doi:
10.1186/1471-2164-13-589
Clarke, K. R., and Gorley, R. N., (2001). Primer v5: User Manual/Tutorial.
PRIMER-E: Plymouth.
Dantas, L. G., Esposito, T., Sousa, A. C. B. D., Félix, L., Amorim, L. L. B., Benko-
Iseppon, A. M., et al. (2015). Low genetic diversity and high dierentiation
among relict populations of the neotropical gymnosperm Podocarpus
sellowii (Klotz.) in the Atlantic Forest. Genetica 143, 21–30. doi: 10.1007/
s10709-014-9809-y
Evanno, G., Regnaut, S., and Goudet, J. (2005). Detecting the number of clusters
of individuals using the soware structure: a simulation study. Mol. Ecol. 14,
2611–2620. doi: 10.1111/j.1365-294X.2005.02553.x
Excoer, L., and Lischer, H. E. L. (2010). Arlequin suite ver 3.5: a new series of
programs to perform population genetics analyses under Linux and Windows.
Mol. Ecol. Resour. 10, 564–567. doi: 10.1111/j.1755-0998.2010.02847.x
Farjon, A., and Filer, D., (2013). An atlas of the world’s conifers: an analysis of their
distribution, biogeography, diversity and conservation status. Leiden: Brill. doi:
10.1163/9789004211810
Frech, C., and Chen, N. (2011). Genome comparison of human and non-human
malaria parasites reveals species subset-specic genes potentially linked to human
disease. PLoS Comput. Biol. 7, e1002320. doi: 10.1371/journal.pcbi.1002320
Frankham, R. (2010). Challenges and opportunities of genetic approaches
to biological conservation. Biol. Conserv. 143, 1919–1927. doi: 10.1016/j.
biocon.2010.05.011
Fu, L. G., Li, N., and Mill, R. R., (1999a). “Taxaceae,” in Flora of China. Eds. Wu,
Z. Y., and Raven, P. H. (Beijing: Science Press and St. Louis, Missouri: Missouri
Botanical Garden Press), 89–98.
Fu, L. G., Li, N., and Mill, R. R., (1999b). “Cephalotaxaceae,” in Flora of China.
Eds. Wu, Z. Y., and Raven, P. H. (Beijing: Science Press and St. Louis, Missouri:
Missouri Botanical Garden Press), 85–88.
Fu, X. H., Deng, S. L., Su, G. H., Zeng, Q. L., and Shi, S. H. (2004). Isolating high-
quality RNA from mangroves without liquid nitrogen. Plant Mol. Biol. Rep. 22,
197. doi: 10.1007/BF02772728
Ge, X. J., Hung, K. H., Ko, Y. Z., Hsu, T. W., Gong, X., Chiang, T. Y., et al.
(2015). Genetic divergence and biogeographical patterns in Amentotaxus
argotaenia species complex. Plant Mol. Biol. Rep. 33, 264–280. doi: 10.1007/
s11105-014-0742-0
Ge, X. J., Zhou, X. L., Li, Z. C., Hsu, T. W., Schaal, B. A., and Chiang, T. Y. (2005).
Low genetic diversity and signicant population structuring in the relict
Amentotaxus argotaenia complex (Taxaceae) based on ISSR ngerprinting. J.
Plant Res. 118, 415–422. doi: 10.1007/s10265-005-0235-1
Geng, Q. F., Liu, J., Sun, L., Liu, H., Ou-Yang, Y., Cai, Y., et al. (2015). Development
and characterization of polymorphic microsatellite markers (SSRs) for an
endemic plant, Pseudolarix amabilis (Nelson) Rehd. (Pinaceae). Molecules 20,
2685–2692. doi: 10.3390/molecules20022685
Hilton-Taylor, C., Yang, Y., Rushforth, K., and Liao, W., (2013). Amentotaxus
argotaenia. e IUCN Red List of reatened Species. Available at: http://www.
iucnredlist.org (Accessed July 3, 2017).
Hoban, S. M., Haue, H. C., Pérez-Espona, S., Arntzen, J. W., Bertorelle, G., Bryja,
J., et al. (2013). Bringing genetic diversity to the forefront of conservation
policy and management. Conserv. Genet. Resour. 5, 593–598. doi: 10.1007/
s12686-013-9859-y
Hu, X. G., Liu, H., Jin, Y., Sun, Y. Q., Li, Y., Zhao, W., et al. (2016). De novo
transcriptome assembly and characterization for the widespread and stress-
tolerant conifer Platycladus orientalis. PLoS ONE 11, e0148985. doi: 10.1371/
journal.pone.0148985
Huang, L. K., Yan, H. D., Zhang, X. X., Zhang, X. Q., Wang, J., Frazier, T.,
et al. (2015). Identifying dierentially expressed genes under heat stress
and developing molecular markers in orchardgrass (Dactylis glomerata
L.) through transcriptome analysis. Mol. Ecol. Resour. 15, 1497–1509. doi:
10.1111/1755-0998.12418
Jump, A. S., and Peñuelas, J. (2005). Running to stand still: adaptation and the
response of plants to rapid climate change. Ecol. Lett. 8, 1010–1020. doi:
10.1111/j.1461-0248.2005.00796.x
Kalinowski, S. T., Taper, M. L., and Marshall, T. C. (2007). Revising how
the computer program CERVUS accommodates genotyping error
increases success in paternity assignment. Mol. Ecol. 16, 1099–1106. doi:
10.1111/j.1365-294X.2007.03089.x
La Rota, M., Kantety, R. V., Yu, J.-K., and Sorrells, M. E. (2005). Nonrandom
distribution and frequencies of genomic and EST-derived microsatellite markers
in rice, wheat, and barley. BMC Genomics 6, 23. doi: 10.1186/1471-2164-6-23
Law, J. A., and Jacobsen, S. E. (2010). Establishing, maintaining and modifying
DNA methylation patterns in plants and animals. Nat. Rev. Genet. 11, 204–220.
doi: 10.1038/nrg2719
Li, C. Y., Chiang, T. Y., Chiang, Y. C., Hsu, H. M., Ge, X. J., Huang, C. C., et al. (2016).
Cross-species, ampliable EST-SSR markers for Amentotaxus species obtained by
next-generation sequencing. Molecules 21, 67. doi: 10.3390/molecules21010067
Lin, C., Chan, M. H., Chen, F. S., and Wang, Y. N. (2007). Age structure and growth
pattern of an endangered species, Amentotaxus formosana Li. J. Integr. Plant
Biol. 49, 157–167. doi: 10.1111/j.1744-7909.2007.00429.x
Liu, L., Zhang, S., and Lian, C. (2015). De novo transcriptome sequencing analysis
of cDNA library and large-scale unigene assembly in Japanese Red Pine (Pinus
densiora). Int. J. Mol. Sci. 16, 29047–29059. doi: 10.3390/ijms161226139
Lu, C. H. (2006). Roles of animals in seed dispersal of Pinus: a review. Chin. J. Ecol.
25, 557–562. doi: 10.1093/rheumatology/kex446
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
Transcriptome of Amentotaxus argotaeniaRuan et al.
10
Nystedt, B., Street, N. R., Wetterbom, A., Zuccolo, A., Lin, Y. C., Scoeld, D. G., et
al. (2013). e Norway spruce genome sequence and conifer genome evolution.
Nature 497, 579–584. doi: 10.1038/nature12211
Ortego, J., Riordan, E. C., Gugger, P. F., and Sork, V. L. (2012). Inuence
of environmental heterogeneity on genetic diversity and structure in
an endemic southern Californian oak. Mol. Ecol. 21, 3210–3223. doi:
10.1111/j.1365-294X.2012.05591.x
Peakall, R., and Smouse, P. E. (2006). GenALEx 6: genetic analysis in Excel.
Population genetic soware for teaching and research. Mol. Ecol. Notes 6, 288–
295. doi: 10.1111/j.1471-8286.2005.01155.x
Pinosio, S., González-Martínez, S. C., Bagnoli, F., C attonaro, F., Grivet, D., Marroni,
F., et al. (2014). First insights into the transcriptome and development of new
genomic tools of a widespread circum-Mediterranean tree species, Pinus
halepensis Mill. Mol. Ecol. Resour. 14, 846–856. doi: 10.1111/1755-0998.12232
Piry, S., Luikart, G., and Cornuet, J. M. (1999). Bottleneck: a computer program
for detecting recent reductions in the eective size using allele frequency data.
J. Hered. 90, 502–503. doi: 10.1093/jhered/90.4.502
Ranade, S. S., Lin, Y. C., Zuccolo, A., Peer, Y. V. D., and García-Gil, M. D. R. (2014).
Comparative in silico analysis of EST-SSRs in angiosperm and gymnosperm
tree genera. BMC Plant Biol. 14, 220. doi: 10.1186/s12870-014-0220-8
Rungis, D., Bérubé, Y., Zhang, J., Ralph, S., Ritland, C. E., Ellis, B. E., et al.
(2004). Robust simple sequence repeat markers for spruce (Picea spp.) from
expressed sequence tags. eor. Appl. Genet. 109, 1283–1294. doi: 10.1007/
s00122-004-1742-5
Sima, Y. K. (2004). Studies of conservation biology on Cephalotaxus oliveri, a rare
medicinal plant. Kunming, China: Yunnan University Press.
Su, Y. J., Wang, T., Zheng, B., Jiang, Y., Chen, G. P., Ouyang, P. Y., et al. (2005).
Genetic dierentiation of relictual populations of Alsophila spinulosa in
southern China inferred from cpDNA trnL–F noncoding sequences. Mol.
Phylogenet. Evol. 34, 323–333. doi: 10.1016/j.ympev.2004.10.016
Ueno, S., Moriguchi, Y., Uchiyama, K., Ujino-Ihara, T., Futamura, N., Sakurai,
T., et al. (2012). A second generation framework for the analysis of
microsatellites in expressed sequence tags and the development of EST-SSR
markers for a conifer, Cryptomeria japonica. BMC Genomics 13, 136. doi:
10.1186/1471-2164-13-136
Ueno, S., Wen, Y., and Tsumura, Y. (2015). Development of EST-SSR markers for
Taxus cuspidata from publicly available transcriptome sequences. Biochem.
Syst. Ecol. 63, 20–26. doi: 10.1016/j.bse.2015.09.016
Van Oosterhout, C., Hutchinson, W. F., Wills, D. P. M., and Shipley, P.
(2004). MICRO-CHECKER: soware for identifying and correcting
genotyping errors in microsatellite data. Mol. Ecol. Notes 4, 535–538. doi:
10.1111/j.1471-8286.2004.00684.x
Wang, C. R., Yan, H. D., Li, J., Zhou, S. F., Liu, T., Zhang, X. Q., et al. (2018a).
Genome survey sequencing of purple elephant grass (Pennisetum purpureum
Schum ‘Zise’) and identication of its SSR markers. Mol. Breed. 38, 94. doi:
10.1007/s11032-018-0849-3
Wang, Y. L., Liang, Q. L., Hao, G. Q., Chen, C. L., and Liu, J. Q. (2018b). Population
genetic analyses of the endangered alpine Sinadoxa corydalifolia (Adoxaceae)
provide insights into future conservation. Biodivers. Conserv. 27, 2275–2291.
doi: 10.1007/s10531-018-1537-7
Wayne, R. K., and Morin, P. A. (2004). Conservation genetics in the new molecular
age. Front. Ecol. Envion. 2, 89–97. doi: 10.1890/1540-9295(2004)002[0089:CGI
TNM]2.0.CO;2
Wegrzyn, J. L., Liechty, J. D., Stevens, K. A., Wu, L. S., Loopstra, C. A., Vasque-
Gross, H. A., et al. (2014). Unique features of the Loblolly Pine (Pinus taeda L.)
megagenome revealed through sequence annotation. Genetics 196, 891–909.
doi: 10.1534/genetics.113.159996
Xiang, X. Y., Zhang, Z. X., Wang, Z. G., Zhang, X. P., and Wu, G. L. (2015).
Transcriptome sequencing and development of EST-SSR markers in Pinus
dabeshanensis, an endangered conifer endemic to China. Mol. Breed. 35, 158.
doi: 10.1007/s11032-015-0351-0
Yu, H., and Li, Q. (2008). Exploiting EST databases for the development and
characterization of EST-SSRs in the Pacic oyster (Crassostrea gigas). J. Hered.
99, 208–214. doi: 10.1093/jhered/esm124
Zalapa, J. E., Cuevas, H., Zhu, H., Stean, S., Senalik, D., Zeldin, E., et al. (2012).
Using next-generation sequencing approaches to isolate simple sequence repeat
(SSR) loci in the plant science. Am. J. Bot. 99, 193–208. doi: 10.3732/ajb.1100394
Zeng, J., Chen, J., Kou, Y. X., and Wang, Y. J. (2018). Application of EST-SSR
markers developed from the transcriptome of Torreya grandis (Taxaceae), a
threatened nut-yielding conifer tree. PeerJ 6, e5606. doi: 10.7717/peerj.5606
Zhang, D. Q., and Zhou, N. (2013). Genetic diversity and population structure of
the endangered conifer Taxus wallichiana var. mairei (Taxaceae) revealed by
Simple Sequence Repeat (SSR) markers. Biochem. Syst. Ecol. 49, 107–114. doi:
10.1016/j.bse.2013.03.030
Zhang, Q., Chiang, T. Y., George, M., Liu, J. Q., and Abbott, R. J. (2005).
Phylogeography of the Qinghai–Tibetan Plateau endemic Juniperus przewalskii
(Cupressaceae) inferred from chloroplast DNA sequence variation. Mol. Ecol.
14, 3513–3524. doi: 10.1111/j.1365-294X.2005.02677.x
Zhou, S. F., Wang, C. R., Yin, G. H., Zhang, Y., Shen, X. Y., Pennerman, K. K.,
etal. (2018a). Phylogenetics and diversity analysis of Pennisetum species using
Hemarthria EST-SSR markers. Grassland Sci. 65, 13–22. doi: 10.1111/grs.12208
Zhou, S. F., Wang, C. R., Frazier, T. P., Yan, H. D., Chen, P. L., Chen, Z. H., et al.,
(2018b). e rst Illumina-based de novo transcriptome analysis and molecular
marker development in Napier grass (Pennisetum purpureum). Mol. Breed. 38,
95. doi: 10.1007/s11032-018-0852-8
Conict of Interest: e authors declare that the research was conducted in the
absence of any commercial or nancial relationships that could be construed as a
potential conict of interest.
Copyright © 2019 Ruan, Wang, Wang and Su. is is an open-access article distributed
under the terms of the Creative Commons Attribution License (CC BY). e use,
distribution or reproduction in other forums is permitted, provided the original
author(s) and the copyright owner(s) are credited and that the original publication in
this journal is cited, in accordance with accepted academic practice. No use, distribution
or reproduction is permitted which does not comply with these terms.
Frontiers in Genetics | www.frontiersin.org October 2019 | Volume 10 | Article 1014
... The quality of a de novo assembly can be assessed by the mean length and N50 value of the contigs. As presented in Table 2, the mean length and N50 value of the unigenes were 1.3 and 2.0 kb, respectively, which were comparatively better than other transcriptomic studies [37][38][39] and other Pistacia transcriptomes published in the literature [40,41]. Although the higher N50 value and greater average length can indicate an accurate and effective assembly [37,42], previous research has suggested that both measures are primitive and often misleading [43]. ...
... Transcriptome sequencing has been commonly used to screen SSRs in various angiosperm species, including Fragaria × Potentilla (red-flowering strawberry), Amentotaxus argotaenia, Curcuma alismatifolia, Vigna angularis, P. vera, and P. chinensis [22,33,[37][38][39][40][41]. In this study, out of 83,370 unigenes, 17,028 unigenes consisted of SSR motifs, accounting for 20.42% of total sequences, with a SSR distribution density of 1 per 5.1 kb (Tables 4 and 5). ...
... The most frequent trinucleotide repeat motifs observed in P. chinensis were in the following order: AAG/CTT (6.75%), AAT/ATT (2.90%), and ACC/GGT (2.61%). Similar findings were also reported in V. angularis [34], B. napus [27], I. polycarpa [36], and A. argotaenia [39]. Taken together, these data indicate that the trinucleotide motif AAG/CTT is common in P. chinensis [22]. ...
Article
Full-text available
Pistacia chinensis Bunge (P. chinensis), a dioecious plant species, has been widely found in China. The female P. chinensis plants are more important than male plants in agricultural production, as their seeds can serve as an ideal feedstock for biodiesel. However, the sex of P. chinensis plants is hard to distinguish during the seedling stage due to the scarcity of available transcriptomic and genomic information. In this work, Illumina paired-end RNA sequencing assay was conducted to unravel the transcriptomic profiles of female and male P. chinensis flower buds. In total, 50,925,088 and 51,470,578 clean reads were obtained from the female and male cDNA libraries, respectively. After quality checks and de novo assembly, a total of 83,370 unigenes with a mean length of 1.3 kb were screened. Overall, 64,539 unigenes (77.48%) could be matched in at least one of the NR, NT, Swiss-Prot, COG, KEGG, and GO databases, 71 of which were putatively related to the floral development of P. chinensis. Additionally, 21,662 simple sequence repeat (SSR) motifs were identified in 17,028 unigenes of P. chinensis, and the mononucleotide motif was the most dominant type of repeats (52.59%) in P. chinensis, followed by dinucleotide (22.29%), trinucleotide (20.15%). The most abundant repeats were AG/CT (13.97%), followed by AAC/GTT (6.75%) and AT/TA (6.10%). Based on these SSR, 983 EST-SSR primers were designed, 151 of which were randomly chosen for validation. Of these validated EST-SSR markers, 25 SSR markers were found to be polymorphic between male and female plants. One SSR marker, namelyPCSSR55, displayed excellent specificity in female plants, which could clearly distinguish between male and female P. chinensis. Altogether, our findings not only reveal that the EST-SSR marker is extremely effective in distinguishing between male and female P. chinensis but also provide a solid framework for sex determination of plant seedlings.
... In this study, a total of 36446 unigenes were used to detect SSRs, and finally 4352 unigenes containing 5089 SSRs were identified. The distribution frequency was 11.94%, which was similar to that of C. hainanensis (11.39%) (Qiao et al., 2014) and P. chienii (11.15%) (Xu et al., 2020), higher than that of P. bungeana (9.21%) (Duan et al., 2017), Torreya grandis (2.75%) (Zeng et al., 2018), A. argotaenia (7.68%) (Ruan et al., 2019), and P. koraiensis (6.84%) (Li et al., 2021), lower than that of Lycium barbarum (27.93%) (Chen et al., 2017), Dalbergia odorifera (23.31%) (Liu et al., 2019), and peony (20.38%) (He et al., 2020). The average density of SSRs was 1/11.1 kb, lower than that of C. hainanensis (1/8.08 kb) (Qiao et al., 2014), Glyptostrobus pensilis (1/7.59 kb) (Li et al., 2019), and P. chienii (1/9.18 kb) (Xu et al., 2020), while higher than that in P. dabeshanensis (1/23.08 kb) (Xiang et al., 2015), P. koraiensis (1/ 17.38 kb) (Du et al., 2017), and L. principis-rupprechtii (1/26.8 ...
... The predominant type was trinucleotide repeats (26.19%), followed by dinucleotide repeats (15.35%) and hexanucleotide repeats (4.30%). This result was similar to that of pervious reports that trinucleotide repeats were the abundant type for other conifers, including C. hainanensis (Qiao et al., 2014), A. argotaenia (Ruan et al., 2019), P. koraiensis , and P. chienii (Xu et al., 2020). Among the trinucleotide repeats, the most abundant motif was AAG/CTT, which was identical to previous findings in Cryptomeria japonica (Ueno et al., 2012), P. halepensis (Pinosio et al., 2014), L. gmelinii , T. grandis (Zeng et al., 2018), A. argotaenia (Ruan et al., 2019), and P. chienii (Xu et al., 2020). ...
... This result was similar to that of pervious reports that trinucleotide repeats were the abundant type for other conifers, including C. hainanensis (Qiao et al., 2014), A. argotaenia (Ruan et al., 2019), P. koraiensis , and P. chienii (Xu et al., 2020). Among the trinucleotide repeats, the most abundant motif was AAG/CTT, which was identical to previous findings in Cryptomeria japonica (Ueno et al., 2012), P. halepensis (Pinosio et al., 2014), L. gmelinii , T. grandis (Zeng et al., 2018), A. argotaenia (Ruan et al., 2019), and P. chienii (Xu et al., 2020). In addition, this motif was the second abundant motif in C. hainanensis (Qiao et al., 2014) and P. dabeshanensis (Xiang et al., 2015). ...
Article
Full-text available
Cephalotaxus oliveri is an endemic conifer of China, which has medicinal and ornamental value. However, the limited molecular markers and genetic information are insufficient for further genetic studies of this species. In this study, we characterized and developed the EST-SSRs from transcriptome sequences for the first time. The results showed that a total of 5089 SSRs were identified from 36446 unigenes with a density of one SSR per 11.1 kb. The most common type was trinucleotide repeats, excluding mononucleotide repeats, followed by dinucleotide repeats. AAG/CTT and AT/AT exhibited the highest frequency in the trinucleotide and dinucleotide repeats, respectively. Of the identified SSRs, 671, 1125, and 1958 SSRs were located in CDS, 3′UTR, and 5′UTR, respectively. Functional annotation showed that the SSR-containing unigenes were involved in growth and development with various biological functions. Among successfully designed primer pairs, 238 primer pairs were randomly selected for amplification and validation of EST-SSR markers and 47 primer pairs were identified as polymorphic. Finally, 28 high-polymorphic primers were used for genetic analysis and revealed a moderate level of genetic diversity. Seven natural C. oliveri sampling sites were divided into two genetic groups. Furthermore, the 28 EST-SSRs had 96.43, 71.43, and 78.57% of transferability rate in Cephalotaxus fortune, Ametotaxus argotaenia, and Pseudotaxus chienii, respectively. These markers developed in this study lay the foundation for further genetic and adaptive evolution studies in C. oliveri and related species.
... Furthermore, there were 89,750 unigenes categorized into three main GO categories: biological processes (233,859; 47.53%); cellular components (145,239; 29.52%); and molecular functions (112,917; 22.95%) (Fig. 3). Within the three categories, 'binding' (53,980), 'cellular process' (52,634), 'metabolic process' (49,492), 'catalytic activity' (41,468), and 'singleorganism process' (37,122) were the most prevalent. Following searches against the KEGG database, 53,059 unigenes were classified into five categories, which were distributed in 130 metabolic pathways. ...
... Recently, 123 SSRs were identified in the A. tsaoko chloroplast genome; the mean density of SSR loci is up to 1/1.33 kb; however, they have not been validated for A. tsaoko [48]. In recent years, the use of transcriptome data to obtain sequences containing microsatellites, and the study of their genetic diversity has been successfully reported [52][53][54]. Excluding mononucleotide repeats, trinucleotides were found to be the most abundant repeats in the present study. ...
Article
Full-text available
Background Amomum tsaoko is a medicinal and food dual-use crop that belongs to the Zingiberaceae family. However, the lack of transcriptomic and genomic information has limited the understanding of the genetic basis of this species. Here, we performed transcriptome sequencing of samples from different A. tsaoko tissues, and identified and characterized the expressed sequence tag-simple sequence repeat (EST-SSR) markers. Results A total of 58,278,226 high-quality clean reads were obtained and de novo assembled to generate 146,911 unigenes with an N50 length of 2002 bp. A total of 128,174 unigenes were successfully annotated by searching seven protein databases, and 496 unigenes were identified as annotated as putative terpenoid biosynthesis-related genes. Furthermore, a total of 55,590 EST-SSR loci were detected, and 42,333 primer pairs were successfully designed. We randomly selected 80 primer pairs to validate their polymorphism in A. tsaoko ; 18 of these primer pairs produced distinct, clear, and reproducible polymorphisms. A total of 98 bands and 96 polymorphic bands were amplified by 18 pairs of EST-SSR primers for the 72 A. tsaoko accessions. The Shannon's information index (I) ranged from 0.477 (AM208) to 1.701 (AM242) with an average of 1.183, and the polymorphism information content (PIC) ranged from 0.223 (AM208) to 0.779 (AM247) with an average of 0.580, indicating that these markers had a high level of polymorphism. Analysis of molecular variance (AMOVA) indicated relatively low genetic differentiation among the six A. tsaoko populations. Cross-species amplification showed that 14 of the 18 EST-SSR primer pairs have transferability between 11 Zingiberaceae species. Conclusions Our study is the first to provide transcriptome data of this important medicinal and edible crop, and these newly developed EST-SSR markers are a very efficient tool for germplasm evaluation, genetic diversity, and molecular marker-assisted selection in A. tsaoko .
... This question is relevant to natural forest regeneration, especially for dioecious species such as D. pectinatum. The initial glimpse into the complexity of the D. pectinatum genome, we found that the GC content of the D. pectinatum transcriptome was 45%, similar to what have been reported in other conifer species, e.g., 47% in Amentotaxus argotaenia [35] and 44.58% in Pinus dabeshanensis [36]. The dominant SSR type was mononucleotide repeats with 9-12 bps. ...
... AP3 orthologs in Pinus tabuliformis and Picea abies, Deficiens-Agamous-Like11 (DAL11), and DAL13 are also male specific [41,42]. Both studies by Fei et al. [35] and Verelst et al. [36] found the expression of AP3 and AGL30 reaching its peak during the meiosis of microsporocyte. In comparison, the D. pectinatum AP3 and AGL30 homologs were found expressed in an earlier stage ( Figure 8). ...
Article
Full-text available
Dacrydium pectinatum de Laubenfels is a perennial gymnosperm species dominant in tropical montane rain forests. Due to severe damages by excessive deforestation, typhoons, and other external forces, the population of the species has been significantly reduced. Furthermore, its natural regeneration is poor. To better understand the male cone development in D. pectinatum, we examined the morphological and anatomical changes, analyzed the endogenous hormone dynamics, and profiled gene expression. The morpho-histological observations suggest that the development of D. pectinatum male cone can be largely divided into four stages: microspore primordium formation (April to May), microspore sac and pollen mother cell formation (July to November), pollen mother cell division (January), and pollen grain formation (February). The levels of gibberellins (GA), auxin (IAA), abscisic Acid (ABA), cytokinin (CTK), and jasmonic acid (JA) fluctuated during the process of male cone development. The first transcriptome database for a Dacrydium species was generated, revealing >70,000 unigene sequences. Differential expression analyses revealed several floral and hormone biosynthesis and signal transduction genes that could be critical for male cone development. Our study provides new insights on the cone development in D. pectinatum and the foundation for male cone induction with hormones and studies of factors contributing to the species’ low rate of seed germination.
... When better characterized and combined with male transcriptome analysis, the unigenes can be applied to revealing the molecular mechanisms of various important biological processes in Dacrydium. In further understanding the characteristics of the D. pectinatum genome, we found that the GC content of the female D. pectinatum transcriptome was 44.99%, the same as the male transcriptome (45%) and in addition, similar to what has been reported in other conifer species, e.g., 47% in Amentotaxus argotaenia (Ruan et al., 2019) and 44.58% in Pinus dabeshanensis (Xiang et al., 2015). The dominant SSR type was mononucleotide repeats with 9-12 bps both in female and male D. pectinatum. ...
Article
Full-text available
Dacrydium pectinatum de Laubenfels is a perennial dioeciously gymnosperm species dominant in tropical montane rain forests. Due to deforestation, natural disasters, long infancy, and poor natural regeneration ability, the population of this species has been significantly reduced and listed as an endangered protected plant. To better understand the female cone development in D. pectinatum, we examined the morphological and anatomical changes, analyzed the endogenous hormone dynamics, and profiled gene expression. The female reproductive structures were first observed in January. The morpho-histological observations suggest that the development of the D. pectinatum megaspore can be largely divided into six stages: early flower bud differentiation, bract primordium differentiation, ovule primordium differentiation, dormancy, ovule maturity, and seed maturity. The levels of gibberellins (GA), auxin (IAA), abscisic acid (ABA), and cytokinin (CTK) fluctuate during the process of female cone development. The female cones of D. pectinatum need to maintain a low level of GA3-IAA-ABA steady state to promote seed germination. The first transcriptome database for female D. pectinatum was generated, revealing 310,621 unigenes. Differential expression analyses revealed several floral (MADS2, AGL62, and LFY) and hormone biosynthesis and signal transduction (CKX, KO, KAO, ABA4, ACO, etc.) genes that could be critical for female cone development. Our study provides new insights into the cone development in D. pectinatum and the foundation for female cone induction with hormones.
... Compared with genomic SSRs (g-SSRs), expressed sequence tag SSRs (EST-SSRs) were developed from cDNA sequences, which located in coding regions and had more cross-species transferability (Cavagnaro et al. 2011;Zhang et al. 2006;Ellis et al. 2007). Currently, a total of 112; 2,156; 1,953; 7,235; 4,955; 5,510; 2,065 potential or putative EST-SSR had been identified from Chamaecyparis formosensis, C. lanceolata, P. densiflora, P. koraiensis, Amentotaxus formosana, Amentotaxus argotaenia, and Torreya grandis transcriptomes, respectively, and some loci were detected to be polymorphic Wen et al. 2015;Liu et al. 2015a;Li et al. 2016Li et al. , 2020cRuan et al. 2019;Zeng et al. 2018). These markers could be used to identify functional genes involved in significant traits, while they could enhance the understanding of genetic diversity, genetic relationship, and population genetics in these species. ...
Article
Full-text available
High-throughput sequencing technology has been widely used and continuously optimized. As one kind of high-throughput sequencing technology, RNA sequencing (RNA-seq) can effectively detect complete transcriptome information, which brings prospects for plant researches. Conifers are a family of Coniferopsida, mainly including Cupressaceae, Taxodiaceae, Pinaceae and Taxaceae trees. Presently, the whole genome sequencing (WGS) of many crops and broadleaf trees have been completed, while that of conifer species are limited due to huge data volume and complicated repeat sequences. Therefore, RNA-seq is one of the most important and practical methods for exploring conifers genes. This review introduced the analysis processes of RNA-seq, and highlighted the applications and prospects of RNA-seq in conifers molecular genetics, which might provide references for future genetic improvement and molecular breeding in conifer species.
... Pinus massoniana (He = 0.5717) (Zhang et al., 2014), Picea abies (He = 0.616) (Stojnić et al., 2019), and P. likiangensis (He = 0.7186) (Cheng et al., 2014); close to P. dabeshanensis (He = 0.36) and Amentotaxus argotaenia (He = 0.39) (Ruan et al., 2019); and higher than A. formosana (He = 0.1993) , P. bungeana (He = 0.205) (Duan et al., 2017), and A. yunnanensis (He = 0.3343) . It has been suggested that levels of genetic variation in conifers are influenced by a variety of factors including lifespan, reproductive system, seed dispersal mechanisms, geographical distribution range, life forms, demographic history, natural selection, and mutation rate (Hamrick et al., 1992;Su et al., 2009;Wang et al., 2020). ...
Article
Full-text available
Pseudotaxus chienii, belonging to the monotypic genus Pseudotaxus (Taxaceae), is a relict conifer endemic to China. Its populations are usually small and patchily distributed, having a low capacity of natural regeneration. To gain a clearer understanding of how landscape variables affect the local adaptation of P. chienii, we applied EST-SSR markers in conjunction with landscape genetics methods: (a) to examine the population genetic pattern and spatial genetic structure; (b) to perform genome scan and selection scan to identify outlier loci and the associated landscape variables; and (c) to model the ecological niche under climate change. As a result, P. chienii was found to have a moderate level of genetic variation and a high level of genetic differentiation. Its populations displayed a significant positive relationship between the genetic and geographical distance (i.e., “isolation by distance” pattern) and a strong fine-scale spatial genetic structure within 2 km. A putatively adaptive locus EMS6 (functionally annotated to cellulose synthase A catalytic subunit 7) was identified, which was found significantly associated with soil Cu, K, and Pb content and the combined effects of temperature and precipitation. Moreover, P. chienii was predicted to experience significant range contractions in future climate change scenarios. Our results highlight the potential of specific soil metal content and climate variables as the driving force of adaptive genetic differentiation in P. chienii. The data would also be useful to develop a conservation action plan for P. chienii.
Article
Full-text available
In this study, the genetic diversity and population structure of 4 wild ancient tea tree (Camellia taliensis) populations at different altitudes (2,050, 2,200, 2,350, and 2,500 m) in Qianjiazhai Nature Reserve, Zhenyuan country, Yunnan province, were investigated using EST-SSR molecular markers to compare their genetic variation against altitude. In total, 182 alleles were detected across all loci, ranging from 6 to 25. The top one informative SSR was CsEMS4 with polymorphism information content (PIC) of 0.96. The genetic diversity of this species was high, with 100% of loci being polymorphic, an average Nei's gene diversity (H) of 0.82, and Shannon's information index (I) of 1.99. By contrast, at the population level, the genetic diversity of wild ancient tea tree was relatively low, with values of H and I being 0.79 and 1.84, respectively. Analysis of molecular variance (AMOVA) revealed a minor genetic differentiation (12.84%) among populations, and most of the genetic variation (87.16%) was detected within populations. Using population structure analysis, we found that the germplasm of wild ancient tea tree was divided into three groups, and there was a substantial gene exchange among these three groups at different altitudes. Divergent habitats caused by altitudes and high gene flow played important roles in genetic diversity of wild ancient tea tree populations, which will provide new opportunities for promoting their protection and potential utilization.
Chapter
In this chapter, the biodiversity, chemodiversity, and pharmacotherapy of Amentotaxus, Pseudotaxus, and Torreya medicinal plants are reviewed. The chloroplast genome sequences of Amentotaxus, Torreya, and Pseudotaxus have been reported. These genera are closely related, and various bioactive compounds, for example, diterpenoid and triterpenoid, flavonoid, lignan, lipid, oil, and trace element, have been isolated from them. Extract and compounds of these genera have antibacterial activity, antiinflammatory activity, cytotoxic and anticancer activity, antioxidant activity, and anthelmintic activity. Toxicity and adverse reactions are also noted. Molecular markers of these genera are very useful in the studies of population genetics, molecular phylogeny, and evolution. Further uses of transcriptomics, proteomics, and metabolomics techniques will facilitate the conservation and sustainable utilization of medicinal resources of these Taxaceae genera.
Article
Full-text available
Expressed sequence tag‐simple sequence repeat (EST‐SSR) markers are widely applied in plant molecular genetics studies due to their abundance in the genome, codominant nature, and high repeatability. To study the genetic diversity of 35 accessions and transferability of EST‐SSR markers in cross‐species applications, 21 primer pairs previously developed in Hemarthria were amplified across Pennisetum species. A total of 148 polymorphisms were generated with an average of 7.05 bands per locus. The mean values of the genetic parameters polymorphism information content, number of bands, effective number of bands, Nei's gene diversity, and Shannon's information index were 0.8430, 2.0000, 1.7640, 0.4247 and 0.6132, respectively. Cluster analysis of 35 accessions divided them into three clusters, which were consistent with dendrograms of 14 species. The among‐population component explained most of the genetic diversity (66%); the remaining variation (34%) occurred within populations. This study will provide more available EST‐SSR markers for Pennisetum species, and facilitate studies on germplasm collection and utilization of Pennisetum species.
Article
Full-text available
Torreya grandis (Taxaceae) is an ancient conifer species endemic to southeast China. Because of its nutrient-rich and delicious seeds, this species has been utilized for centuries by the Chinese. However, transcriptome data and transcriptome-derived microsatellite markers for population genetics studies are still insufficient for understanding of this species’ genetic basis. In this study, a transcriptome from T. grandis leaves was generated using Illumina sequencing. A total of 69,920 unigenes were generated after de novo assembly, and annotated by searching against seven protein databases. In addition, 2,065 expressed sequence tag–simple sequence repeats (EST-SSRs) were detected, with the distribution frequency of 2.75% of total unigenes and average number of 0.03 SSRs per unigene. Among these EST-SSRs, 1,339 primer pairs were successfully designed, and 106 primer pairs were randomly selected for the development of potential molecular markers. Among them, 11 EST-SSR markers revealed a moderate level of genetic diversity, and were used to investigate the population structure of T. grandis . Two different genetic groups within this species were revealed using these EST-SSR markers, indicating that these markers developed in this study can be effectively applied to the population genetic analysis of T. grandis .
Article
Full-text available
Pennisetum purpureum belongs to the Pennisetum Rich genus in the family Poaceae. It is widely grown in subtropical and tropical regions as one of the most economically important cereal crops. Despite its importance, there is limited genomic data available for P. purpureum, which restricts genetic and breeding studies in this species. In the present study, the transcriptome of P. purpureum was assembled de novo and used to characterize two important P. purpureum cultivars: P. purpureum Schumab cv. Purple and P. purpureum cv. Mott. After assembly, a total of 197,466 unigenes were obtained for ‘Purple’ and ‘Mott’ and 103,454 of these unigenes were successfully annotated. From ‘Purple’ and ‘Mott,’ 214,648 SNPs and 21,213 EST-SSRs were identified in 40,259 unigenes and 18,587 unigenes respectively. Moreover, 50 EST-SSR primers and 6 SNP primers were designed to validate the identified markers. The transcriptomic data of present study from the two P. purpureum cultivars provides an abundant amount of available genomic information for Pennisetum. In addition, the identified SNPs and EST-SSRs will facilitate genetic and molecular studies within the Pennisetum genus.
Article
Full-text available
Elephant grass (Pennisetum purpureum) is a perennial grass in the Poaceae family with high tolerance and one of the best forage plants. Despite its economic importance, the inheritance information of P. purpureum has remained largely unknown. To obtain the whole reference genome, we first conducted a genome survey of P. purpureum. Next-generation sequencing (NGS) was used to perform the de novo whole genome sequencing. As a result, the estimated genome size of elephant grass was 2.01 Gb, with 71.36% repetitive elements. The heterozygosity was 1.02%, which indicates a highly heterozygous genome. The retroelements (9.36%) were the most repetitive elements, followed by DNA transposons (3.66%). In the meantime, 83,706 high-quality genomic simple sequence repeat (SSR) markers, in which the greatest SSR unit length was 3, were developed. Thirty pairs of SSR markers were randomly selected to verify the efficiency and all of them yielded clear amplification products, among which 28 pairs (93.3%) of the primers showed polymorphism. The genome data obtained in this research provided a large amount of gene resources for further investigating Pennisetum species.
Article
Full-text available
With increasing temperature and anthropogenic activity, endangered alpine species in the high altitudes of the Qinghai-Tibet Plateau face high risk of extinction; however, they have received little attention in the past. In this study, we used 12 nuclear and nine chloroplast microsatellites (simple sequence repeats, SSR) to assess genetic diversity within and among the only two populations of the highly endangered alpine species Sinadoxa corydalifolia (Adoxaceae). We identified only one individual exhibiting clonal reproduction across all 160 extant plants. The levels of genetic variability were estimated to be very low, with the allele number Na = 3.2 and the expected heterozygosity He = 0.368. The genetic differentiation is extremely high between the two regional populations (FST = 0.214), with a limited rate of gene flow in the recent past. In addition, numerous endemic alleles were found for each subpopulation within each population. Our analyses suggest that it is critical not only to conserve all surviving individuals of the two populations in situ but also to mediate gene flow artificially between subpopulations within each population in this endangered species.
Article
Full-text available
Platycladus orientalis, of the family Cupressaceae, is a widespread conifer throughout China and is extensively used for ecological reforestation, horticulture, and in medicine. Transcriptome assemblies are required for this ecologically important conifer for understanding genes underpinning adaptation and complex traits for breeding programs. To enrich the species’ genomic resources, a de novo transcriptome sequencing was performed using Illumina paired-end sequencing. In total, 104,073,506 high quality sequence reads (approximately 10.3 Gbp) were obtained, which were assembled into 228,948 transcripts and 148,867 unigenes that were longer than 200 nt. Quality assessment using CEGMA showed that the transcriptomes obtained were mostly complete for highly conserved core eukaryotic genes. Based on similarity searches with known proteins, 62,938 (42.28% of all unigenes), 42,158 (28.32%), and 23,179 (15.57%) had homologs in the Nr, GO, and KOG databases, 25,625 (17.21%) unigenes were mapped to 322 pathways by BLASTX comparison against the KEGG database and 1,941 unigenes involved in environmental signaling and stress response were identified. We also identified 43 putative terpene synthase (TPS) functional genes loci and compared them with TPSs from other species. Additionally, 5,296 simple sequence repeats (SSRs) were identified in 4,715 unigenes, which were assigned to 142 motif types. This is the first report of a complete transcriptome analysis of P. orientalis. These resources provide a foundation for further studies of adaptation mechanisms and molecular-based breeding programs.
Article
Full-text available
Amentotaxus, a genus of Taxaceae, is an ancient lineage with six relic and endangered species. Four Amentotaxus species, namely A. argotaenia, A. formosana, A. yunnanensis, and A. poilanei, are considered a species complex because of their morphological similarities. Small populations of these species are allopatrically distributed in Asian forests. However, only a few codominant markers have been developed and applied to study population genetic structure of these endangered species. In this study, we developed and characterized polymorphic expressed sequence tag-simple sequence repeats (EST-SSRs) from the transcriptome of A. formosana. We identified 4955 putative EST-SSRs from 68,281 unigenes as potential molecular markers. Twenty-six EST-SSRs were selected for estimating polymorphism and transferability among Amentotaxus species, of which 23 EST-SSRs were polymorphic within Amentotaxus species. Among these, the number of alleles ranged from 1-4, the polymorphism information content ranged from 0.000-0.692, and the observed and expected heterozygosity were 0.000-1.000 and 0.080-0.740, respectively. Population genetic structure analyses confirmed that A. argotaenia and A. formosana were separate species and A. yunnanensis and A. poilanei were the same species. These novel EST-SSRs can facilitate further population genetic structure research of Amentotaxus species.
Article
Full-text available
Japanese red pine (Pinus densiflora) is extensively cultivated in Japan, Korea, China, and Russia and is harvested for timber, pulpwood, garden, and paper markets. However, genetic information and molecular markers were very scarce for this species. In this study, over 51 million sequencing clean reads from P. densiflora mRNA were produced using Illumina paired-end sequencing technology. It yielded 83,913 unigenes with a mean length of 751 bp, of which 54,530 (64.98%) unigenes showed similarity to sequences in the NCBI database. Among which the best matches in the NCBI Nr database were Picea sitchensis (41.60%), Amborella trichopoda (9.83%), and Pinus taeda (4.15%). A total of 1953 putative microsatellites were identified in 1784 unigenes using MISA (MicroSAtellite) software, of which the tri-nucleotide repeats were most abundant (50.18%) and 629 EST-SSR (expressed sequence tag- simple sequence repeats) primer pairs were successfully designed. Among 20 EST-SSR primer pairs randomly chosen, 17 markers yielded amplification products of the expected size in P. densiflora. Our results will provide a valuable resource for gene-function analysis, germplasm identification, molecular marker-assisted breeding and resistance-related gene(s) mapping for pine for P. densiflora.
Book
An Atlas of the World's Conifers is the first ever atlas of all known conifer species. It is based on locality information of ca. 37,000 collected herbarium specimens held in scientific institutions. As well as providing natural distribution maps for each species, Farjon and Filer give the reader comprehensive insight into the biogeography, diversity and conservation status of conifers on all continents, dispelling the widely held view that they are primarily a northern boreal plant group. Conifer diversity is analysed and presented with a taxonomic and geographic perspective. Distribution patterns are interpreted using the latest information on continental drift, dispersal and phylogeny. The entire dataset supporting the Atlas can be consulted and verified online. These data can also be used for further research and are an invaluable resource for anyone working on conifer systematics, biogeography or conservation.