Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Untangling Nucleotide Diversity and Evolution of the H Genome in Polyploid Hordeum and Elymus Species Based on the Single Copy of Nuclear Gene DMC1

  • Dongfa Sun,

    Affiliation College of Plant Science and Technology, Huazhong Agricultural University, Wuhan, People's Republic of China

  • Genlou Sun

    genlou.sun@smu.ca

    Affiliations College of Plant Science and Technology, Huazhong Agricultural University, Wuhan, People's Republic of China, Biology Department, Saint Mary's University, Halifax, Nova Scotia, Canada

Abstract

Numerous hybrid and polypoid species are found within the Triticeae. It has been suggested that the H subgenome of allopolyploid Elymus (wheatgrass) species originated from diploid Hordeum (barley) species, but the role of hybridization between polyploid Elymus and Hordeum has not been studied. It is not clear whether gene flow across polyploid Hordeum and Elymus species has occurred following polyploid speciation. Answering these questions will provide new insights into the formation of these polyploid species, and the potential role of gene flow among polyploid species during polyploid evolution. In order to address these questions, disrupted meiotic cDNA1 (DMC1) data from the allopolyploid StH Elymus are analyzed together with diploid and polyploid Hordeum species. Phylogenetic analysis revealed that the H copies of DMC1 sequence in some Elymus are very close to the H copies of DMC1 sequence in some polyploid Hordeum species, indicating either that the H genome in theses Elymus and polyploid Hordeum species originated from same diploid donor or that gene flow has occurred among them. Our analysis also suggested that the H genomes in Elymus species originated from limited gene pool, while H genomes in Hordeum polyploids have originated from broad gene pools. Nucleotide diversity (π) of the DMC1 sequences on H genome from polyploid species (π = 0.02083 in Elymus, π = 0.01680 in polyploid Hordeum) is higher than that in diploid Hordeum (π = 0.01488). The estimates of Tajima's D were significantly departure from the equilibrium neutral model at this locus in diploid Hordeum species (P<0.05), suggesting an excess of rare variants in diploid species which may not contribute to the origination of polyploids. Nucleotide diversity (π) of the DMC1 sequences in Elymus polyploid species (π = 0.02083) is higher than that in polyploid Hordeum (π = 0.01680), suggesting that the degree of relationships between two parents of a polyploid might be a factor affecting nucleotide diversity in allopolyploids.

Introduction

Hybridization and polyploidization have played a central role in the history of plant evolution, and contributed greatly to plant diversification and speciation [1], [2]. Much attention has been drawn to studying the evolutionary consequences of polyploid species in both genome size and contents, with the advances in molecular methods over the last two decades [3], [4]. Polyploid genome origins and evolution have also been the focus of plant evolutionists [1], [5]. Increasing evidences have demonstrated the complexity of the dynamic nature of polyploids. Many polyploids are proved to involve multiple origins in space and time [1], [5], together with introgression (or gene flow) [6][8]. Both multiple origins [9], [10] and gene flow [6][8] have been considered as the causes of shared polymorphism across ploidy level and/or phylogenetic incongruence among loci. However, whether gene flow among independent formations is regular occurrence following polyploid species have rarely been tested in ployploid taxa [11].

The tribe Triticeae contains several important cereal crops such as wheat, barley and rye, as well as forage crops. The tribe combines a wide variety of biological mechanisms and genetic systems which makes it an excellent group for research in evolution, genetic diversity, taxonomy, and speciation in plants [12]. According to Löve [13] and Dewey's [14] classification, genus Hordeum and Elymus are two relative large genera in the tribe Triticeae.

The genus Hordeum comprises 31 species (including cultivated barley, H. vulgare ssp. vulgare) and exists at the diploid, tetraploid, and hexaploid levels with a basic chromosome number x = 7. Based on cytogenetic analyses, the diploid species in Hordeum were classified into four monogenomic groups: H, I, Xa, and Xu genome group [15], [16], which were supported by isoenzyme analysis [17] and molecular data [18][22]. The H genome group is not only the largest genome group in Hordeum (including 14 diploid species, 7 tetraploid species, 4 hexaploid species, and 2 species existing at three ploidy levels (2×, 4×, 6×), and distributed widely from central Asia to the Americas), but also widely present in polyploid species in Elymus, Stenostachys and Pascopyrum.

Within the genus Elymus are approximately 50 allotetraploid species that combined both H and St genomes, and distributed throughout the world in non-tropical areas, from northern Greenland in the north to Tierra del Fuego in southernmost South America [23]. The St haplome originated from the genus Pseudoroegneria [14]. It has been confirmed that the H haplomes in Elymus were contributed by different Hordeum diploids [24][30]. Phylogenetic analyses based on phosphoenolpyruvate carboxylase, β-amylase, granule-bound starch synthase I [29] and disrupted meiotic cDNA (DMC) [30], suggested few potential Hordeum diploid species as H-genome donors to Elymus species. The tetraploid H. jubatum might have involved in the origin of StH Elymus [29]. However, the role of polyploid Hordeum species in the origin of StH Elymus remains to be studied. It is not clear whether gene flow across polyploid Hordeum and Elymus species has occurred following polyploid speciation. Recent studies led to the conclusion that the polyploid probably originated multiple times [1], [5], which are often considered as a potential source of increased genetic variation in polyploids. However, how much genetic variation is contributed by the diploid progenitors and the degree of gene flow among the independent origins are the two major factors determining the genetic diversity in polyploids. Yet, the extent and role of gene flow among polyploids in evolution remains enigmatic.

In present study, DMC1 data from the allopolyploid StH Elymus are analyzed together with diploid and polyploid Hordeum species. The objectives of this analysis are: (1) to explore the possible role of polyploid Hordeum species in the origin of StH Elymus; (2) to determine whether gene flow has occurred between polyploid Hordeum and Elymus; and (3) to examine the level of nucleotide polymorphism in the H genomes from Elymus, Hordeum diploids and polyploids. Answering these questions will provide new insights into the formation of these polyploid species, and the potential role of gene flow among polyploid species during polyploid evolution.

Materials and Methods

Samples

The present study includes 18 tetraploid (22 accessions) StH Elymus, 9 polyploid Hordeum species. All diploid Hordeum species and other diploid Triticeae species representing the St, W, P, and E genomes were included for analyses (Table 1). Bromus arvensis and B. sterilis were used as outgroups. The single copy nuclear gene disrupted meiotic cDNA (DMC1) has been applied to phylogenetic analyses in Triticeae species. The DMC1 sequences used in this study were collected from previously published sources [21], [30][35].

Data Analysis

Multiple sequence alignments are made using Clustal X with default parameters [36] with manual adjustment. Phylogenetic analysis using the maximum-parsimony (MP) method is performed with the computer program PAUP* ver. 4 beta 10 [37]. All characters are specified as unweighted and unordered, and gaps are excluded in the analyses. Most-parsimonious trees are obtained by performing a heuristic search using the Tree Bisection-Reconnection (TBR) option with MulTrees on, and ten replications of random addition sequences with the stepwise addition option. Multiple parsimonious trees are combined to form a strict consensus tree. Overall character congruence is estimated by the consistency index (CI), and the retention index (RI). In order to infer the robustness of clades, bootstrap values with 1000 replications [38] are calculated by performing a heuristic search using the TBR option with MulTrees on.

In addition to maximum parsimony analysis, maximum-likelihood (ML) analysis is performed. For ML analysis, 8 nested models of sequence evolution were tested for the data set using PhyML 3.0 [39]. The general time-reversible (GTR) [40] substitution model led to a largest ML score compared to the other 7 substitution models: JC69, K80, F81, F84, HKY85, TN93 and custom (data not shown). As the result, the GTR model was used for the ML analysis. The ML analysis was performed using the Mac OS X UNIX version of GARLI v. 0.95 [41]. The runs were set for an unlimited number of generations, and automatic termination following 10,000 generations without a significant (lnL increase of 0.01) topology change. Thirty analyses were run with random starting tree topologies, and the tree with best score was used to represent gene tree. Branch support (BS) was estimated based on 100 ML bootstrap replicates in GARLI.

Nucleotide diversity was estimated by Tajima's π [42] and Watterson's θ [43] statistics. The former measure quantifies the mean percentage of nucleotide differences among all pairwise comparisons for a set of sequences, whereas the latter is simply an index of the number of segregating (polymorphic) sites. Tests of neutral evolution were performed as described by Tajima [42], and Fu and Li [44]. The above calculations were conducted by the software program DnaSP v5 [45].

Results

Total of 87 sequences from 18 tetraploid (22 accessions) StH Elymus, 9 polyploid Hordeum species, 24 diploid Hordeum species/subspecies and 8 other diploid Triticeae species were analyzed. Sequence comparisons revealed five large insertions/deletions (indels). Compared to other sequences aligned here, one copy sequence (H1166s4) from hexaploid H. procerum and the sequence from diploid H. cordobense (AY134715) showed a 24 bp insertion at position 206. One sequences from E. confusus (PI 598463k) showed a 23 bp deletion at position 352. An 82 bp insertion was found in the sequences from polyploid species H. fuegianum, H. jubatum, and H. tetraploidum, and diploid Australopyrum velutinum and Taeniatherum caput-medusae as reported by Wang and Sun [35], and Petersen and Seberg [31], respectively. The sequence from E. trachycaulus (PI 537323L) showed a 30 bp insertion at position 1054. A 15 bp deletion was detected in the sequences from E. canadensis (EU366405), E. cordilleranus (H6486k), E. hystrix (EU366415), H. arizonicum (H2144s3), H. brachyantherum subsp. californicum (AF277260), H. brachyantherum subsp. brachyantherum (H2348s2), H. depressum (H2008s1) and H. procerum (H1166s3).

The 89 (including two outgroups) aligned 1221 bp DMC sequences showed 794 constant, 221 variable and parsimony-uninformative, and 206 parsimony-informative sites. Parsimony analysis using Bromus arvensis and B. sterilis as outgroup produced 740 equally parsimonious trees with a consistency index (CI) of 0.693, and a retention index (RI) of 0.848. Maximum likelihood analysis across 30 GARLI runs generated likelihood score ranging from –lnL6349.08703 to –lnL6355.83219. ML tree with BS is shown in Figure 1.

thumbnail
Figure 1. The best scoring ML tree was selected from 30 GARLI analyses under a GTR model.

Numbers above branches are ML bootstrap support values, and the numbers below branches are bootstrap support values.

https://doi.org/10.1371/journal.pone.0050369.g001

Two copies of sequences each from E. caninus, E. cordilleranus, E. hystrix, E. sibiricus, E. virginicus and E. wawawaiensis were well separated into two distinct groups, one grouped with the sequences from H genome, and another with St genome from Pseudoroegneria (Fig. 1). As unexpected, the sequence (GQ855194) from E. transhyrcanus formed clade with Lophopyrum elongatum and Thinopyrum bessarabicum with 90% BS in ML and 78% BS support in MP. The second copy of the sequence from E. transhyrcanus (GQ855193) was sister to the sequences from E. canadensis, E. sibiricus and E. multisetus. The sequences from polyploid Hordeum species fall into the clade with sequences from diploid Hordeum except one each from polyploid H. arizonicum, H. fuegianum, H. jubatum, and H. tetraploidum (Fig. 1). Also included in this clade are the sequences from Elymus species which were from the H genome. One sequence from H. arizonicum was clustered with the St genome sequences from E. cordilleranus, E. hystrix, E. canadensis, and E. wiegandii. An Asian diploid H. brevisubulatum subsp. violaceum is sister to polyploid H. lechleri, H. parodii, H. procerum, H. tetraploidum, E. lanceolatus, E. confusus, E. virescens, E. wawawaiensis, E. caninus. Within this clade, the sequence from E. lanceolatus is sister to the sequence from H. lechleri, H. parodii, H. procerum, H. tetraploidum with highly support (87% BS in ML, 85% in MP). One copy each from polyploid H. arizonicum, H. brachyantherum subsp. brachyantherum, H. depressum, H. procerum, E. cordilleranus, E. hystrix and American diploid species H. brachyantherum subsp. californicum formed a moderately supported clade (83% BS in ML, and 62% BS in MP) (Fig. 1).

Based on grouping of the sequences in phylogenetic analysis, we further separately analyzed nucleotide variation of DMC1 gene in the H genome from Hordeum polyploids and diploids, and Elymus. Some of the putative H copies of sequences from Elymus and Hordeum polyploids were not clearly put into the H clade. These sequences (PI 537323L, GQ855194 and W6-13828K from Elymus, H6198s2, H2013s2, H1418s2 and H2144s2 from Hordeum) were excluded for nucleotide diversity analysis. Estimates of nucleotide polymorphism, π and θw, were shown separately for the H genome of Elymus, Hordeum polyploid and diploid species (Table 2). The number of polymorphic sites (56) in the H genome of polyploid Hordeum is much lower than that (90) in its diploid donor species. The estimates of nucleotide diversity in the H genome of diploid Hordeum studied were θw = 0.02693, π = 0.01488. The estimates of nucleotide diversity in the H genome of polyploid Hordeum studied were θw = 0.0168, π = 0.01734. The number of polymorphic sites, the estimates of nucleotide diversity θw, π for the H genome in Elymus species was 80, 0.02774 and 0.02083, respectively. The Tajima [42], and Fu & Li's [44] tests were conducted on each data set. The Tajima, and Fu and Li's values of the diploid Hordeum H genome were −1.95371 (P<0.05) and −2.65118 (P<0.05), respectively, which showed significant departure from neutrality, while the Tajima, and Fu and Li's values of the polyploid Hordeum H genome were −0.34466 and −0.80485, respectively. The Tajima's D was −1.19959 and the Fu and Li's D was −1.29141 for Elymus H genome.

thumbnail
Table 2. Estimates of nucleotide diversity and test statistics for selection at DMC1 in polyploidy and diploid H genome.

https://doi.org/10.1371/journal.pone.0050369.t002

Discussion

Origin of H genome in StH-genome Elymus species based on single copy nuclear gene DMC1 has previously been discussed [30]. DMC1 sequence data also showed a reticulate relationship of American polyploid species and diploid Hordeum [35]. However, the relationship of the H genome in polyploid Hordeum and Elymus was not previously explored.

The maximum parsimonious analysis grouped 24 sequences from Hordeum diploid and polyploid species together with 94% bootstrap supported value, maximum likelihood analysis also grouped these sequences together with highly supported value of 97% (Fig. 1). Only 3 Hordeum diploid H genome species, H. brevisubulatum subsp. violaceum, H. brachyantherum subsp. californicum, and H. bogdanii were grouped together with the sequences from Elymus H genome, indicating that the H genomes in Elymus originated from limited Hordeum diploid species, whereas the H genomes in polyploid Hordeum species were contributed by relative large Hordeum diploids. One concern is that the number of sequences from Elymus H genome is less than the number of sequences from Hordeum polyploids analyzed here, which may bias the comparison. However, phylogeny of Elymus StStHH allotetraploids based on three nuclear genes including relative large sample of Elymus species suggested that the one diploid Hordeum species, H. brachyantherum subsp. californicum (Syn: H. californicum Covas & Stebbins), is the possible H- genome donor to Elymus species [29], which also indicated that H genome in Elymus species originated from limited Hordeum diploid. However, the published data indicated that many Hordeum diploid species have contributed to the origin of polyploids in this genus, more than 10 diploid species were suggested to be the potential donors to polyploids in Hordeum [22], [35], [46]. Taken these together, it can be concluded that the H genomes in Elymus species have originated from limited gene pool, while H genomes in Hordeum polyploids have originated from broad gene pools.

Polyploid formation is a prominent mode of speciation in the flowering plant. Recent molecular data indicated that polyploid speciation is often more complex than initially thought [47], which is also the case in the tribe Triticeae [8], [22], [28], [29], [48], [49]. Molecular studies suggested polyploid species in many genera have multiple origins rather than single origin [1], [5], [47], [50]. It was suggested that the fates of polyploid populations of independent origins varied depending on the amount of genetic variation initially contributed by the diploid progenitors [50]. Studies have demonstrated that genetic diversity in polyploids is often similar to or higher than their diploid progenitors [47], [51], [52]. It is worth comparing the nucleotide diversities among the H genomes from Elymus, polyploid and diploid Hordeum species. This may offer an opportunity to address the potential evolutionary outcomes of polyploidization. Nucleotide diversity (π) of the DMC1 sequences from polyploid species (π = 0.02083 in Elymus, π = 0.01680 in polyploid Hordeum) is higher than that in diploid Hordeum (π = 0.01488). The estimates of D were significantly departure from the equilibrium neutral model at this locus in diploid Hordeum species (P<0.05), suggesting an excess of rare variants in diploid species. These rare variants may not contribute to the origination of polyploids. Phylogenetic analyses indeed indicated that not all diploids have contributed to the origination of polyploids in Hordeum and Elymus. Why is the genetic variation in polyploids higher than in diploid even the polyploids originated from limited number of diploid detected here? It has demonstrated that polyploidization resulted in the genome wide gene duplication which not only enables allopolyploids to tolerate more genomic variation than their progenitors, but also provides new opportunity to create functional diversification between homoeologous genes over time [52][54]. After gene duplicated, one of the copies may undergo mutations, if mutations are not deleterious, the mutations will not be removed by natural selection. Nucleotide diversity of this copy of gene in polyploids will be higher than that in their progenitors. Recent studies on the evolutionary rates of duplicated genes in polyploids compared to their diploid relatives showed that the evolutionary rates appear to be different among different homoeologous locus pairs [55][58]. Barrier et al. [59] found that the floral regulatory genes APETALA1 (ASAP1) and APETALA3 (APETALA3/TM6) are evolving much faster in the polyploid species than in the diploids. Analysis of nucleotide sequence diversity (π) of RPB2 revealed that nucleotide diversity (π) of RPB2 on the St genome in tetraploid Elymus was higher than that in the diploid Pseudoroegneria St genome [60].

The degree of relationships between two parents of a polyploid was suggested as a general factor affecting the amount of genomic sequence variation in allopolyploids [54]. In a study on interspecifc crosses of Brassica found that the overall amount of genomic change in AC (or CA) tetraploids was much lower than that in the AB (or BA) tetraploids. This was because the genetic distance between the A (B. rapa) and C (B. oleracea) was much closer than that between the A and B (B. nigra) [61]. A study on the timing and rate of genome variation in triticale following allopolyploidization also suggested the degree of the relationship between the parental genomes was the key factor in determining the rate of genomic sequences variation occurring during intergeneric allopolyploidization [54]. It was well demonstrated that genus Hordeum is monophyletic genus [21], [22], and polyploids originated from the diploid species in this genus. While the Elymus StH genomic species originated from the St genome donor Pseudoroegneria and H genome donor Hordeum species. The genetic distance of parental genomes in polyploid Hordeum is much closer than that between St and H genomes. The degree of relationships between two parents of a polyploid might be factor affecting nucleotide diversity in allopolyploids. This might explain that nucleotide diversity (π) of the DMC1 sequences in Elymus polyploid species (π = 0.02083) is higher than that in polyploid Hordeum (π = 0.01680). This speculation needs to be further studied.

One of objectives of this study is to explore the possible role of polyploid Hordeum species in the origin of StH Elymus and whether the gene flow has occurred between polyploid Hordeum and Elymus species. The Hordeum H genome diploids are the H genome donor to both Elymus StH species and polyploid species in Hordeum. Phylogenetic analysis revealed that the H copy of DMC1 sequence in E. lanceolatus is very close to the H copy of DMC1 sequence in polyploid H. procerum, H. lechleri, H. parodii, and H. tetraploidum, indicating that the H genome in E. lanceolatus and those four polyploid Hordeum species originated from the same diploid donor. Alternative explanation is that gene flow has occurred among E. lanceolatus and H. procerum, H. lechleri, H. parodii, and H. tetraploidum. Elymus cordilleranus and E. hystrix formed a group with ployploid H. arizonicum, H. brachyantherum subsp. brachyantherum, H. depressum, H. procerum, and diploid H. brachyantherum subsp. californicum. This grouping is explained either by common origin of H genome in these Elymus and Hordeum polyploids or by gene flow among these polyploids. One copy of the DMC1 sequences from North American hexaploid H. arizonicum was grouped with St-copy sequences from American E. canadensis, E. hystrix, E. wiegandii and E. cordilleranus (Fig. 1). The diploid H. pusillum and tetraploid H. jubatum was suggested as the parental parents for H. arizonicum [22], [46], [62]. cpDNA analysis suggested that H. pusillum could be the maternal parent of H. arizonicum [63]. Previous analysis of DMC1 data suggested that H. brachyantherum subsp. californicum might be one diploid genome donor of H. arizonicum, and the second genome donor of H. arizonicum might be the common ancestor of H. brachyantherum subsp. brachyantherum, and showed that one copy of DMC1 sequences of H. arizonicum fall outside the Hordeum clade of the tree [35]. DMC1 data here suggested if the St genome was not the donor species to one copy of genome in H. arizonicum, gene flow has occurred between H. arizonicum and some Elymus StH genome species during evolutionary process.

Analysis of β-amylase data revealed that one of the H. jubatum genome was placed together with Elymus species [29]. The role of H. jubatum in the Elymus evolutionary history has been suggested, a tetraploid similar to H. jubatum might have been involved in the history of Elymus, either through introgression between the Elymus and H. jubatum, or through a direct contribution from H. jubatum like species to Elyums [29]. Our DMC1 data showed that one of the H. jubatum genome with H. tetraploidum, H. fuegianum, T. caput-medusae and Aust. velutinum grouped together, Elymus virescens as sister to this group. Our result not only did not contradict to the suggestion that H. jubatum was involved at some stage in the history of StStHH Elymus [29], but also further expanded to that several ployploid Hordeum species might have involved in the evolution of StStHH Elymus through gene flow among them.

The study on the polyploid formation in Tragopogon (Asteraceae) indicated a lack of gene flow among polyploid plants of independent origin, even when they co-occur, suggesting potential reproductive barriers among separate lineages in polyploid species [50]. Sequence analysis of 12 nuclear loci representing 6 genes on tetraploid Capsella bursa-pastoris and its close diploid relative C. rubella showed that polyploid speciation need not result in immediate and complete reproductive isolation, and the postpolyploidization hybridization and introgression can contribute significantly to genetic variation in newly formed polyploid [64]. Molecular characterization of a diagnostic DNA marker for domesticated tetraploid wheat indicated gene flow from wild tetraploid wheat to hexaploid wheat [65]. Our results suggested that gene flow among different polyploids in Triticeae species might play an important role in polyploid speciation and evolution.

Acknowledgments

We thank two anonymous reviewers for critical comments and suggestions on the manuscript which significantly improved the manuscript.

Author Contributions

Conceived and designed the experiments: GS DS. Performed the experiments: GS. Analyzed the data: GS DS. Contributed reagents/materials/analysis tools: GS DS. Wrote the paper: DS GS.

References

  1. 1. Soltis DE, Soltis PS (1999) Polyploidy: Recurrent formation and genome evolution. Trends Ecol Evol 14: 348–352.
  2. 2. Cui L, Wall PK, Leebens-Mack JH, Lindsay BG, Soltis DE, et al. (2006) Widespread genome duplications throughout the history of flowering plants. Genome Res 16: 738–749.
  3. 3. Osborn TC, Pires JC, Birchler JA, Auger DL, Chen ZJ, et al. (2003) Understanding mechanisms of novel gene expression in polyploids. Trends Genet 19: 141–147.
  4. 4. Wendel JF (2000) Genome evolution in polyploids. Plant Mol Biol 42: 225–249.
  5. 5. Soltis DE, Soltis PS, Tate JA (2003) Advances in the study of polyploidy since plant speciation. New Phytolog 161: 173–191.
  6. 6. Lihová J, Shimizu KK, Marhold K (2006) Allopolyploid origin of Cardamine asarifolia (Brassicaceae): Incongruence between plastid and nuclear ribosomal DNA sequences solved by a single-copy nuclear gene. Mol Phylogen Evol 39: 759–786.
  7. 7. Mason-Gamer RJ (2004) Reticulate evolution, introgression, and intertribal gene capture in an allohexaploid grass. Syst Biol 53(1): 25–37.
  8. 8. Mason-Gamer RJ (2008) Allohexaploidy, introgression, and the complex phylogenetic history of Elymus repens (Poaceae). Mol Phylogen Evol 47: 598–611.
  9. 9. Trewick SA, Morgan-Richards M, Russell SJ, Henderson S, Rumsey FJ, et al. (2002) Polyploidy, phylogeography and Pleistocene refugia of the rockfern Asplenium ceterach: evidence from chloroplast DNA. Mol Ecol 11: 2003–2012.
  10. 10. Vanichanon A, Blake NK, Sherman JD, Talbert LE (2003) Multiple origins of allopolyploid Aegilops triuncialis. Theor Appl Genet 106: 804–855.
  11. 11. Espinoza NR, Noor MA (2002) Popuplationgenetics of a polyploidy: is there hybridization between lineages of Hyla versicolor? J Hered 93: 81–85.
  12. 12. Bothmer von R, Salomon B (1994) Triticeae: a tribe for food, feed and fun. In: Wang RR –C, Jensen KB, Jaussi C, editors. Proc. 2nd Int. Triticeae Symp., Utah. Utah State University Publication Design and Production, Logan, pp. 1–12.
  13. 13. Löve A (1984) Conspectus of the Triticeae. Feddes Report 95: 425–521.
  14. 14. Dewey DR (1984) The genomic system of classification. A guide to intergeneric hybridization with the perennial Triticeae. In: Gustafson JP, editor. Gene manipulation in Plant Improvement. New York: Plenum Press. pp. 209–280.
  15. 15. Bothmer von R, Fink J, Landstrom T (1986) Meiosis in interspecific Hordeum hybrids. I. Diploid combinations. Can J Genet Cytol 28: 525–535.
  16. 16. Bothmer von R, Fink J, Landstrom T (1987) Meiosis in interspecific Hordeum hybrids. II. Triploid hybrids. Evol Trends Plants 1: 42–50.
  17. 17. Jørgensen RB (1986) Relationships in the barley genus (Hordeum): an electrophoretic examination of proteins. Hereditas 104: 165–221.
  18. 18. Baum BR, Bailey LG (1991) Relationships among native and introduced North American species of Hordeum, based on chloroplast DNA restriction site variation. Can J Bot 69: 2421–2426.
  19. 19. Doebley J, von Bothmer R, Larson S (1992) Chloroplast DNA variation and the phylogeny of Hordeum (Poaceae). Am J Bot 79: 576–584.
  20. 20. Svitashev S, Bryngelsson T, Vershinin A, Pedersen C, Säll T, et al. (1994) Phylegenetic analysis of the genus Hordeum using repetitive DNA sequences. Theor Appl Genet 89: 801–810.
  21. 21. Petersen G, Seberg O (2003) Phylogenetic analyses of the diploid species of Hordeum (Poaceae) and a revised classification of the genus. Syst Bot 28: 293–306.
  22. 22. Blattner FR (2004) Phylogenetic analysis of Hordeum (Poaceae) as inferred by nuclear rDNA ITS sequences. Mol Phylogenet Evol 33: 289–299.
  23. 23. Sun GL, Salomon B (2009) Molecular evolution and origin of the tetraploid Elymus species. Breed Sci 59: 487–491.
  24. 24. Jaaska V (1992) Isoenzyme variation in the grass genus Elymus (Poaceae). Hereditas 117: 11–22.
  25. 25. Wang RR-C, Hsiao C (1986) Differentiation of H genomes of the genus Critesion: Evidence from synthetic hybrids involving Elymus and Critesion and one natural hybrid of C. violaceum and C. bogdanii. Can J Genet Cytol 28: 947–953.
  26. 26. Linde-Laursen I, Seberg O, Salomon B (1994) Comparison of the Giemsa C-banded and N-banded karyotypes of two Elymus species, E. dentatus and E. glaucescens (Poaceae: Triticeae). Plant Syst Evol 192: 165–176.
  27. 27. Sun GL, Salomon B, Bothmer von R (1997) Analysis of tetraploid Elymus species using wheat microsatellite markers and RAPD markers. Genome 40: 806–814.
  28. 28. Sun GL, Ni Y, Daley T (2008) Molecular phylogeny of RPB2 gene reveals multiple origin, geographic differentiation of H genome, and the relationship of the Y genome to other genomes in Elymus species. Mol Phylogenet Evol 46: 897–907.
  29. 29. Mason-Gamer RJ, Burns MM, Naum M (2010) Reticulate evolutionary history of a complex group of grasses: phylogeny of Elymus StStHH allotetraploids based on three nuclear genes. PloS ONE 5(6): e10989 .
  30. 30. Sun GL, Zhang XD (2011) Origin of the H genome in StH-genomic Elymus species based on the single-copy nuclear gene DMC1. Genome 54: 655–662.
  31. 31. Petersen G, Seberg O (2000) Phylogenetic evidence for excision of Stowaway miniature inverted-repeat transposable elements in Triticeae (Poaceae). Mol Biol Evol 17: 1589–1596.
  32. 32. Petersen G, Seberg O (2004) On the origin of the tetraploid species Hordeum capense and H. secalinum (Poaceae). Syst Bot 29: 862–873.
  33. 33. Petersen G, Seberg O, Yde M, Berthelsen K (2006) Phylogenetic relationships of Triticum and Aegilops and evidence for the origin of the A, B, and D genomes of common wheat (Triticum aestivum). Mol Phylogenet Evol 39: 70–82.
  34. 34. Sha L, Fan X, Yang R, Kang H, Ding C, et al. (2010) Phylogenetic relationships between Hystrix and its closely related genera (Triticeae: Poaceae) based on nuclear Acc1, DMC1 and chloroplast trnL-F sequences. Mol Phylogenet Evol 54: 327–335.
  35. 35. Wang H, Sun GL (2011) Molecular phylogeny and reticulate origins of several American polyploid Hordeum species. Botany 89: 405–415.
  36. 36. Thompson JD, Gibson TJ, Plewniak F, Jeanmouguin F, Higgins DG (1997) The CLUSTAL X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucl Acids Res 25: 4876–4882.
  37. 37. Swofford DL (2003) PAUP. Phylogenetic Analysis using Parsimony, version 4. Sunderland, MAUSA: Sinaeur Associates.
  38. 38. Felsenstein J (1985) Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39: 783–791.
  39. 39. Guindon S, Gascuel O (2003) A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol 52: 696–704.
  40. 40. Lanave C, Preparata G, Saccone C, Serio G (1984) A new method for calculating evolutionary substitution rates. J Mol Evol 20: 86–93.
  41. 41. Zwickl DJ (2006) Genetic algorithm approaches for the phylogenetic analysid of large biological sequence datasets under the maximum likelihood criterion. Ph.D dissertation, The University of Texas at Austin
  42. 42. Tajima F (1989) Statistical method for testing the neutral mutation of hypothesis by DNA polymorphism. Genetics 123: 585–595.
  43. 43. Watterson GA (1975) On the number of segregation sites in genetical models without recombination. Theor Popul Biol 7: 256–276.
  44. 44. Fu Y–X, Li W–H (1993) Statistical tests of neutrality of mutations. Genetics 133: 693–709.
  45. 45. Librado P, Rozas J (2009) DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 25: 1451–1452 .
  46. 46. Taketa S, Ando H, Takeda K, Ichii M, von Bothmer R (2005) Ancestry of American polyploid Hordeum species with the I genome inferred from 5S and 18S–25S rDNA. Ann Bot 96: 23–33.
  47. 47. Soltis DE, Soltis PS (1993) Molecular data and the dynamic nature of polyploidy. Crit Rev Plant Sci 12: 243–273.
  48. 48. Mahelka V, Kopecky D (2010) Gene capture from across the grass family in the allohexaploid Elymus repens (L.) Gould (Poaceae: Triticeae) as evidenced by ITS, GBSSI, and molecular cytogenetics. Mol Biol Evol 27: 1370–1390.
  49. 49. Sun GL, Komatsuda T (2010) Origin of the Y genome in Elymus and its relationship to other genomes in Triticeae based on evidence from elongation factor G (EF-G) gene sequences. Mol Phylogenet Evol 56: 727–733.
  50. 50. Symonds VV, Soltis PS, Soltis DE (2010) Dynamics of polyploidy formation in Tragopogon (Asteraceae): recurrent formation, gene flow, and population structure. Evolution 64: 1984–2003.
  51. 51. Brochmann C, Soltis DE, Soltis PS (1992) Electrophoretic relationships and phylogeny of Nordic polyploids in Draba (Brassicaceae). Plant Syst Evol 182: 35–70.
  52. 52. Otto SP, Whitton J (2000) Polyploid incidence and evolution. Ann Rev Genet 34: 401–437.
  53. 53. Adams KL, Wendel JF (2005) Polyploidy and genome evolution in plants. Curr Opin Plant Biol 8: 135–141.
  54. 54. Ma XF, Gustafson JP (2006) Timing and rate of genome variation in triticale following allopolyploidization. Genome 49: 950–958.
  55. 55. Small RL, Ryburn JA, Wendel JF (1999) Low levels of nucleotide diversity at homoeologous Adh loci in allotetraploid cotton (Gossypium L.). Mol Biol Evol 16: 491–501.
  56. 56. Small RL, Wendel JF (2002) Differential evolutionary dynamics of duplicated paralogous Adh loci in allotetraploid cotton (Gossypium). Mol Biol Evol 19: 597–607.
  57. 57. Lawton-Rauh A, Robichaux RH, Purugganan MD (2003) Patterns of nucleotide variation in homoeologous regulatory genes in the allotetraploid Hawaiian silversword alliance (Asteraceae). Mol Ecol 12: 1301–1313.
  58. 58. Caldwell KS, Dvorak J, Lagudah ES, Akhunov E, Luo MC, et al. (2004) Sequence polymorphism in polyploid wheat and their D-genome diploid ancestor. Genetics 167: 941–947.
  59. 59. Barrier M, Robichaux RH, Purugganan MD (2001) Accelerated regulatory gene evolution in an adaptive radiation. Proc Natl Acad Sci USA 98: 10208–10213.
  60. 60. Sun GL, Daley T, Ni Y (2007) Molecular evolution and genome divergence at RPB2 gene of the St and H genome in Elymus species. Plant Mol Biol 64: 645–665.
  61. 61. Song K, Lu P, Tang K, Osborn TC (1995) Rapid genome change in synthetic polyploids of Brassica and its implications for polyploidy evolution. Proc Natl Acad Sci USA 92: 7719–7723.
  62. 62. Rajhathy T, Symko S (1966) The synthesis of a species: Hordeum arizonicum. Can J Bot 44: 1224–1228.
  63. 63. Nishikawa T, Salomon B, Komatsuda T, von Bothmer R, Kadowaki K (2002) Molecular phylogeny of the genus Hordeum using three chloroplast DNA sequences. Genome 45: 1157–1166.
  64. 64. Slotte T, Huang H, Lascoux M, Ceplitis A (2008) Polyploid speciation did not confer instant reproductive isolation in Capsella (Brassicaceae). Mol Biol Evol 25: 1472–1481.
  65. 65. Dvorak J, Akhunov ED, Akhunov AR, Deal KR, Luo MC (2006) Molecular characterization of a diagnostic DNA marker for domesticated tetraploid wheat provides evidence for gene flow from wild tetraploid wheat to hexaploid wheat. Mol Biol Evol 23: 1386–1396.