Open Access
How to translate text using browser tools
1 December 2014 The Role of Allotriploidy in the Evolution of Meconopsis (Papaveraceae): A Preliminary Study of Ancient Polyploid and Hybrid Speciation
Wei Xiao, Beryl B. Simpson
Author Affiliations +
Abstract

Also known as the Himalayan Poppy, Meconopsis is a genus of herbaceous plants growing only in the high elevation habits of the Himalaya and its adjacent plateau and mountain areas. The genus exhibits high morphological and ecological diversity, but the major causes of divergence in Meconopsis have not previously been studied. Our recent revised taxonomic classification, based on a molecular phylogeny, divided the genus into four monophyletic sections. Because chromosome number varies among these sections and our previous phylogenetic analyses revealed extensive incongruence between the recovered nrITS and cpDNA trees, possibly due to ancient hybridization, this study focused on evaluating the potential role of ancient polyploidization and hybridization in Meconopsis' evolutionary history. Our investigation based on the results of reconstructed ancestral chromosome numbers using a Maximum Likelihood method implemented in chromEvol showed that two extant Meconopsis sections (sect. Grandes and sect. Primulinae) shared a triploid ancestor. We further examined the pattern of hybridization in Meconopsis by reconstructing a nuclear marker GAPDH (glyceraldehyde 3-phosphate dehydrogenase gene) network. The result, along with morphological, phylogenetic, and cytological evidence, all point to a hybrid nature of the triploid ancestor. Based on the resultant GAPDH network, an ancient reticulate evolution scenario in Meconopsis is proposed. Overall, this preliminary study shows how an ancient triploid event promoted polyploid evolution in Meconopsis and also exemplifies how allotriploidization and successive polyploidization played an important role in diversification of the genus.

Meconopsis is a genus distributed across the high elevation of the Himalaya, the southeast Tibetan Plateau, and the Hengduan Mountains. The genus contains ca. 50 species, and is among the most popular and desirable horticultural groups in British gardens. Although there have been efforts to cultivate most Meconopsis species, only a minority of them are able to be grown successfully in gardens. In addition, there has also been extensive crossbreeding for new cultivars (Cobb, 1984). Our phylogenetic study (Xiao, 2013) provided a well-resolved chloroplast (cpDNA) phylogeny for the genus, which led to the recognition and recircumscription of four sections (Fig. 1). Although there has been substantial work on morphology, phylogenetics, and cytology, there is still no coherent framework that synthesizes all the present knowledge into a comprehensive evolutionary explanation of Meconopsis diversity. Because various high polyploid levels exist in Meconopsis (top right Fig. 1), polyploidization may have had an important impact on the evolution of the genus. A recent report documented a synthesized Meconopsis polyploidy (neopolyploid) in Scotland (McNaughton, 2014) which further highlighted the potential role of polyploid evolution. Our interest, therefore, is to understand how different ploidy levels evolved in Meconopsis and to decipher if there has been polyploid speciation in the genus. Unlike neopolyploids, however, since ancient polyploidy formation cannot be experimentally re-created, we took the approach of reconstructing ancestral chromosome numbers on a molecular phylogenetic tree.

Fig. 1.

Overviews of Meconopsis sections. Chloroplast phylogeny is shown on top right (Xiao, 2013), along with the traditional taxonomic classifications from Prain (1915) and Taylor (1934).

i1097-993X-17-1-5-f01.tif

The second aim of this study was to investigate if ancient hybridization significantly contributed to the early divergence of Meconopsis. In our revised classification (Xiao, 2013), the four sections of Meconopsis are monophyletic with each exhibiting consistency in morphology and known chromosome numbers. However, the evolutionary relationships between sections indicated by the cpDNA tree largely conflicted with those inferred from previous morphology-based taxonomic classifications (Prain, 1915; Taylor, 1934). A brief overview of how Meconopsis was previously treated based on morphology is shown in Fig. 1 (top right). Prain's (1915) sections were based on indumentum variations while Taylor's (1934) subsectional treatment was based on whether the leaf rosette is persistent through winter or not. Comparing these two treatments, it is clear that M. sect. Grandes is the “controversial” group in the genus. The Grandes clade was grouped with sect. Meconopsis in Prain's (1915) work, but with M. sect. Aculeatae and M. sect. Primulinae species in Taylor's (1934) treatment. It thus appears that morphological affinity of sect. Grandes to neither M. sect. Meconopsis nor M. sect. Aculeatae is strongly supported. On the other hand, species in M. sect. Primulinae had always previously been grouped or mixed together with M. sect. Aculeatae species taxonomically, an arrangement clearly contradicted by our cpDNA phylogenetic topology. Our phylogenetic study (Xiao, 2013) also revealed incongruence between the nrITS and the cpDNA trees at deep nodes, suggesting that ancient hybridizations may have occurred. These morphological and molecular data led us to the question whether ancient hybridizations were involved in the origin of M. sect. Primulinae and M. sect. Grandes. In order to explore this possibility, we constructed a phylogeny using low-copy nuclear genes with no recombination, or a low recombination rate, to recover the phylogenetic signals from both the parents if reticulation occurred.

Material and Methods

Reconstructing polyploid evolution and ancestral chromosome numbers

Phylogeny reconstruction We firstly calculated a cpDNA tree to serve as the phylogenetic framework for ancestral chromosome number reconstruction. Because Meconopsis is closely related to Papaver and Roemeria (Xiao, 1913), we included Papaver and Roemeria species with Cathcartia as outgroups, in order to achieve an estimation of rates of dysploidy formation and/or genome duplication rates. This phylogenetic reconstruction employed the concatenated sequences (trnL-trnF, matK, ndhF and rbcL) of Meconopsis and Cathcartia species, obtained from our phylogenetic study of Meconopsis (Xiao, 2013) (sequence information in Appendix 1); and 27 previously published trnL-trnF sequences downloaded from Genbank (sequence information is listed in Appendix 2), mainly including Papaver and Roemeria species. Sequences were assembled in Geneious 5.5 (Biomatters, New Zealand), and aligned by Geneious Alignment with the default settings. Absent markers were treated as missing characters. Alignments were then refined manually. Bayesian partition analysis was applied using MrBayes 3.1.2 (Huelsenbeck & Ronquist, 2001), with each marker treated as a partition. The evolutionary models were selected using jModelTest (Posada, 2008). We applied the models most similar to the best-fit models estimated by jModelTest that were available in MrBayes v3.1.2 for each partition: GTR+G for rbcL, GTR+I+G for ndhF, GTR+G for matK, and GTR+I+G for trnL-trnF dataset. Prior probability distributions on all parameters were set to the defaults. Stationarity was reached at twenty million generations using a Markov chain Monte Carlo (MCMC) method. Trees were collected every 100th generation. With 25% burn in, a 50% majority-rule consensus tree was calculated.

Previous cytological studies (Ratter, 1968; Kadereit, 1987; Lavania & Srivastava, 1999; Ying et al., 2006; Kumar et al., 2013) have provided chromosome numbers for many species. For estimating ancestral characters, parsimony based methods have been widely adopted but parsimony approaches usually suffer from strong bias due to their inability to incorporate multiple models and to deal with uncertainty. Mayrose et al. (2010) developed a likelihood method implemented in chromEvol, which estimates the probability of a given ploidy level at any internal node in a given tree. In chromEvol, eight different models are available for testing, each representing a different hypothesis by estimating a different set of parameters (Table 1). For example, the model CONST_RATE_DEMI is characterized by parameters of a constant rate for gaining a single chromosome (gainCONSTR), a constant rate for losing a single chromosome (lossCONSTR), and a constant rate for both whole genome duplication (duplConstR) and demi-polyploidization (demiPloidyR). This model thus hypothesizes that dysploidy, whole genome duplication, and demi-polyploidization (change from 2n to 3n, e.g., forming a triploid) all have occurred along a phylogeny at a constant rate. In contrast to the CONST_RATE_DEMI model, some other models do not allow whole genome duplication or demi-polyploidization. Some of these estimate linear rates (instead of constant rates) to implement the hypothesis that chromosome number change rate is dependent on the current chromosome number. Four of these models were also used. Detailed model comparisons can be retrieved from  http://www.tau.ac.il/∼itaymay/cp/chromEvol/index.html. All the eight models were run to select for the best-fit model. The selected model was used to estimate ancestral chromosome numbers.

Table 1.

Model test result in chromEvol (sorted by AIC score).

i1097-993X-17-1-5-t01.tif

Reconstructing a GAPDH phylogenetic network to detect ancient reticulation

Amplification, PCR, cloning and sequencing – We screened the EST library of Papaver somniferum and related species to choose genes with a single or low-copy number. Among the candidate genes, GAPDH (partial) was selected for its relatively low copy number (we obtained only one copy of the GAPDH gene for the accession of Papaver alpinum, a diploid taxon), successful PCR amplification, and the possession of four introns (five exons) that are phylogenetically informative. Genomic DNA was extracted from silica-dried leaf materials or herbarium materials using the DNeasy Plant Minikit (Qiagen, Valencia, California, USA). Primers were designed with the reference to the EST sequences of Papaver somiferum and other species in Papaveraceae using the EST database from GenBank. The primer pair, forward primer 5′-CACCACCAACTGTCTTGCTCCCCT-3′ and reverse primer 5′-AGCACCCACACTGAAGAGGGAC-3′, were selected. PCR amplification conditions were optimized and carried out in 25 µL reaction volumes with 20–40 ng DNA, 1.0 unit of Taq polymerase (made by the author), 0.5X Failsafe Buffer B (Epicentre Biotechnologies, Madison, WI, USA), and 2.0 µmol/L primers. Eighteen PCR cycles were performed at 95° C for 30 seconds, 59°C for 40 seconds, and 72° C for 45 seconds for each cycle; followed by 26 cycles of 95° C for 30 seconds, 55°C for 40 seconds, and 72° C for 45 seconds. PCR products were visualized on agarose gel containing Syber Safe DNA gel stain (Invitrogen, Eugene, Oregon, USA). Successfully amplified products were cloned using the TOPO-TA cloning kit (Invitrogen, Carlsbad, CA, USA). Thirty to 60 colonies were picked from each plate and amplified by M13 plasmid primers from the cloning kit with the manufacturers' protocols. Successfully amplified products were cleaned using ExoSap (Exonuclease I: New England Biolabs Beverly, MA, USA; Shrimp Alkaline Phosphatase: Progema, Madison, WI, USA) with the manufacturers' protocols. Cleaned PCR products were sequenced on an ABI 3730 DNA Analyzer at the Institute for Cell and Molecular Biology Core Facility at The University of Texas at Austin. GAPDH sequence information is listed in Appendix 3.

Test for recombinant sequences – A potential difficulty of using nuclear markers is that if Meconopsis were a highly polyploid group, multiple copies of the target gene would be amplified simultaneously. Under these circumstances, artificial recombinants could be introduced during PCR when multiple gene copies coexist. We applied an assumption when screening out PCR-mediated recombination that recombinants can only occur once in the raw sequence data but sequences of true genes can reoccur. This assumption is based on the mechanism of PCR-mediated recombination that interruption of complete extension (for example, Taq polymerase can stop extension when mismatching occurs) during PCR produces an incomplete sequence, which anneals to its homologous gene strand and continues to elongate. We consider these “interruptions of extension” as random events which are unlikely to be repeated. We further tested the accuracy of our method by amplifying and examining fragments (< 600 bp) of the GAPDH genes, using our designed internal primers (F1_5′-TATTTTCAATCATTTGTTTC-3′, R1_5′-AATCATTGCAT CCGAGAACAA-3′; F2_5′-AACAGTTTAGTTGCCAATTCG-3′, R2_5′-CTCAATAC TGAAAATTTTG CTAG-3′). We applied this test to all of the Meconopsis species in this study and all of results confirmed that our method performed accurately.

After excluding PCR recombinants, we used the program Recombination Detection Program (Version RDP2) (Martin et al., 2005; available from  http://darwin.uvigo.es/rdp/ rdp.html) to exam natural recombination. We used the analysis algorithms RDP, GENECONV, Bootscan/Recscan, MaxChi, Chimaera, SiScan, and 3seq, which are implemented in RDP2. Detected natural recombinants were eliminated from further analysis.

Alignment and phylogenetic analysis – Alignments were performed using Geneious (Biomatters, New Zealand) Alignment 5.5 with the default settings, and then refined manually. The GAPDH sequences that lost multiple introns were eliminated from the final alignment. Phylogenetic analysis was carried out by RAxML 7.2.8 (Stamatakis, 2006; Stamatakis et al., 2008) with 1000 bootstrap replications.

Rooting of the tree – We obtained only one copy of GAPDH gene from the outgroup species Papaver alpinum which belongs to Papaver section Meconella, the sister group to Meconopsis. However, this sequence was difficult to align with Meconopsis sequences due to a great deal of sequence divergence and phylogenetic analyses using tentative alignments including this outgroup sequence could not resolve the relationship between the major clades with certainty (bootstrap value < 60 for the node directly shared by the outgroup and some Meconopsis species in unrooted trees). Therefore, we chose to use mid-point rooting.

Results

Reconstructing polyploid evolution and ancestral chromosome number

The reconstructed cpDNA Bayesian tree of Meconopsis, Roemeria, and Papaver, rooted by Cathcartia, is shown in Fig. 2A, which is consistent with the Papaver phylogeny published by Carolan et al. (2006). The partial tree that contains only Meconopsis species is expanded in Fig. 2B.

Fig. 2A.

The cpDNA Bayesian trees of Meconopsis, Roemeriana, and Papaver rooted using Cathcartia with chromosome numbers of extant species at the branch tips. Ancestral chromosome numbers reconstructed using chromEvol are shown with the most probable chromosome number (2n with p value > 0.50) labeled above the tree branches and the posterior probabilities below the branches.

Fig. 2B. Reconstructed ancestral chromosome numbers for Meconopsis are shown above the branches and their posterior probabilities below the branches. Known chromosome numbers of extant species are shown at branch tips. Each section is labeled with a different color, corresponding to Fig. 3.

i1097-993X-17-1-5-f02.tif

Fig. 3.

GAPDH phylogenetic network reconstruction in Meconopsis. Bootstrap value labeled above the branch when greater than 50. The three different patterns on the branches (left side) that lead to each of the clade (Clades 1-3) are used to symbolize the sequence origin for each clade and also corresponding to those in Fig. 4. Four colors indicate four sections: each of the colored lines (on the right) connects different copies of the same accession, revealing hybridization pattern; colored dots indicate the accessions with only one obtained GAPDH sequence.

i1097-993X-17-1-5-f03.tif

Fig. 4.

Summary of ancient reticulate hypotheses in Meconopsis. (Note: this graph was transformed from Fig. 3. For example, sect. Aculeatae sequences have multiple origins from both Clades 1 & 3, represented by solid grey-patterned branch and dotted-line branch, respectively, in both Figs. 3 & 4).

i1097-993X-17-1-5-f04.tif

Of the eight models implemented in chromEvol, the four that allowed DEMI (i.e., allowing chromosome number transition from 2n to 3n) all resulted in significantly higher likelihood scores than the other four models without DEMI setting (Table 1). The four DEMI models all generated the same ancestral chromosome numbers, which are shown in Fig. 2A and Fig. 2B. In Fig. 2A, the ancestral chromosome number of the genus Meconopsis is 2n = 14 (posterior probability 0.99). The ancestral chromosome numbers for each section, as shown in Fig. 2B, are 2n = 21 for M. sect. Grandes, 2n = 22 for M. sect. Primulinae, 2n = 28 for M. sect. Meconopsis, 2n = 14 for M. sect. Aculeatae. The chromosome number of the most recent common ancestor shared by M. sect. Grandes and M. sect. Primulinae was estimated to be 2n = 21.

Reconstructing GAPDH phylogenetic network

We successfully obtained partial GAPDH sequences from 21 Meconopsis accessions and one outgroup accession of Papaver alpinum. Because M. sect. Primulinae species are distributed narrowly and endemically, there were few good quality samples for successful PCR amplification. Thus, M. sect. Primulinae is less well represented in the GAPDH network than other sections. On average, 44% of the raw sequences from one PCR reaction were identified as PCR-mediated recombinants. One natural recombination in the sample X022 was detected by program RDP2. After the removal of the recombined sequences, we obtained an average of 2.1 copies of GAPDH gene per accession (1.3 in M. sect. Meconopsis; 2.3 in M. sect. Grandes; 2.5 in M. sect. Aculeatae; 2.0 in M. sect. Primulinae) to reconstruct the network. The average sequence length was 1403 bp with 885 variable sites and an intron/exon ratio of 3.4/1. The ML tree with best likelihood score is shown in Fig. 3 with the bootstrap value labeled above the branch when greater than 50. Three well supported major clades (labeled Clade 1, 2, and 3) are indicated for the ease of discussion (Fig. 3) and each is highlighted by a uniquely patterned branch (left in Fig. 3). Clade 1 comprises only species in M. sect. Aculeatae; Clade 2 contains species of M. sect. Meconopsis, M. sect. Grandes, and M. sect Primulinae; and Clade 3 includes sequences from M. sect. Aculeatae, M. sect. Grandes and M. sect. Primulinae. The network shows that GAPDH sequences in each M. sect. Aculeatae, M. sect. Grandes, and M. sect. Primulinae have multiple origins. We consequently transformed Fig. 3 to the simplified reticulate structure shown in Fig. 4 in order to illustrate our proposed scenario of ancient reticulate evolution in Meconopsis. The patterned branches in Fig. 4 correspond to those in Fig. 3.

Discussion

Hybridization and Allotripolyploid evolution in Meconopsis

According to the ancestral chromosome reconstruction (Fig. 2B), there was an ancestor, with 2n = 21 (marked by the ★ in Fig. 2B) that gave rise to two extant ploidy groups in Meconopsis, M. sect. Grandes and M. sect. Primulinae. In the phylogenetic network using the low-copy nuclear marker GAPDH (Fig. 3 in which each Meconopsis section is indicated by a different color) the GAPDH sequences are clustered into three groups labeled Clades 1, 2, and 3. It is obvious that sequences of M. sect. Grandes and M. sect. Primulinae have multiple origins, located in both Clade 2 and Clade 3 on the GAPDH network (Fig. 3), which strongly disagrees with the cpDNA tree topology where each section is monophyletic (Fig. 2B). One explanation for this pattern of disagreement is that M. sect. Grandes and M. sect. Primulinae had hybrid origin(s). Under this hypothesis, Clade 2, similar to the cpDNA tree (Fig. 2B), could contain the maternal lineage and Clade 3 could represent the paternal lineage for both M. sect. Primulinae and M. sect. Grandes (Fig. 3). In addition, M. sect. Primulinae is morphologically most similar to M. sect. Aculeatae according to traditional classifications of Meconopsis (Prain, 1915; Taylor, 1934) but it is most distant from M. sect. Aculeatae on the cpDNA tree. A hybrid origin could also explain this morphological divergence pattern as hybrid offspring can sometimes predominantly display the traits of one of its parents.

Given our finding that Meconopsis sect. Primulinae and M. sect. Grandes shared a common ancestor of 2n = 21 (★ in Fig. 2B), the formation of this triploid ancestor was probably the result of a hybridization event, which fits the phylogenetic pattern displayed in Fig. 3. Alternatively, it is possible to hypothesize that two or more independent ancient hybridizations in M. sect. Primulinae and M. sect. Grandes respectively could also account for the network structure shown in Fig. 3. However, because, as shown in Fig. 2B, the ancestral chromosome numbers in M. sect. Meconopsis and M. sect. Aculeatae are (2n) 14, 28 or 56, we think hybridization between these cytological types with any of the deviants from a triploid (2n = 21) is unlikely.

It is also possible that the structure of GAPDH network could also be explained by gene duplication and gene loss. However, we considered this a less likely scenario because it would have to require multiple gene-loss events in the early history of the genus to result in the pattern of the observed GAPDH network. Also, the hybridization hypothesis fits the incongruence between nrITS tree and cpDNA tree observed from our phylogenetic study of Meconopsis (Xiao, 2013). However, this alternative hypothesis should be tested further in the future.

We therefore reiterate that the most likely scenario is that Meconopsis sect. Grandes and M. sect. Primulinae shared a triploid ancestor of 2n = 21 with a 2n = 14 maternal parent (evident when tracing back along the cpDNA phylogeny in Fig. 2B). This triploid ancestor (★ in Fig. 2B) appeared to be transient and later established stable lineages (with even-numbered chromosomes) through different putative pathways as follows:

i1097-993X-17-1-5-e01.gif
i1097-993X-17-1-5-e02.gif
i1097-993X-17-1-5-e03.gif

Thus our results suggest that the emergence of two Meconopsis sections was attributable to a single ancient triploidization event. However, mechanisms of how this triploid hybrid ancestor successfully bypassed the low fertility and reproductive instability associated with aneuploidy and hybrid sterility (Comai, 2005) are unclear. It has been suggested that some reproductive strategies, e.g., apomixis, or being perennial, can prolong the life cycle and could contribute to higher chances of establishing reproductively stable lineages (Ramsey & Schemske, 1998). Presumably because a polycarpic life cycle (longevity) can increase chances of the occurrence of a mutation that might lead to eventual reproductive success, the polycarpic habit is frequently found associated with allopolyploidy while diploid populations of the same species are predominantly monocarpic (Treier et al., 2009). In Meconopsis, the majority of species are strictly monocarpic; polycarpic plants are only found in Meconopsis sect. Primulinae (i.e., M. bella) and M. sect. Grandes (i.e., M. betonicifolia, M. grandis). The long-lived polycarpic habit might be derived from, or related to, an allotriploid origin.

According to McNaughton (2014), fertility was gained over time in Meconopsis ‘Lingholm’, a cultivar produced by human hybridization between two species with different ploidy levels – apparently followed the sequence of steps given below:

i1097-993X-17-1-5-e04.gif

McNaughton's report (2014) documented the formation of a new hexaploid (M. ‘Lingholm’) via a sterile triploid (M. × sheldonii), whose descendant, a few decades later, started to produce viable seeds with double the chromosome numbers of its progenitor. McNaughton (2014) commented that Meconopsis triploid cultivars were not uncommon in gardens but that the conversion of a triploid into a fertile hexaploid, as from M. × sheldonii to M. ‘Lingholm’, is a unique event, thanks to which, gardeners can now grow their own “blue heaven” (M. ‘Lingholm’). Our results that recovered an ancient triploid viewed in light of the evidence from M. ‘Lingholm’, suggest that polyploidy was a significant evolutionary force in Meconopsis and that polyploidization through triploidy may have occurred more than once through Meconopsis history. However, the mechanisms (e.g., somatic doubling, fusion of unreduced gametes) of a triploid's transition to a stable hexaploid or dodecaploid (as in M. sect. Grandes) a still largely remain hypothetical (Kadereit, 1987; Bretagnolle & Thompson, 1995; Ramsey & Schemske, 1998; Rieseberg & Willis, 2007; McNaughton, 2014).

In another relevant study, Kadereit (1986, 1987) presented the plausibility of a triploid origin in Papaver somniferum (2n = 20, 22, 44). He proposed a triploid origin of P. somniferum by providing several different lines of evidence supporting that P. somniferum was derived from a triploid hybrid (2n = 21). A later molecular study (Nessler, 1994) analyzing the MLP (major latex proteins) gene family supported an hypothesis of a triploid hybrid origin of the opium poppy. According to Kadereit (1991), the triploid overcame low fertility and stabilized as even number ploidy through regular bivalent formation. He (1987) also suggested that 2n = 22 is an indication of how a triploid (2n = 21) establish even-numbered ploids in Papaver and its closely related genera based on the chromosome number records of a few unrelated species sharing the same aneuploid number n = 11: Meconopsis bella (2n = 22), Papaver somniferum ssp. somniferum (2n = 20, 2n = 22), Papaver somniferum ssp. setigrum (2n = 44), Papaver aculeatum (2n = 22), Roemeria hybrida (2n = 22), Papaver cambricum (2n = 14, 22, 28). The high likelihood scores among the different models calculated by chromEvol (Table 1) with the DEMI parameters (allowing the genome transition from 2n to 3n) support Kadereit's (1986, 1987) conclusion, suggesting that these processes are probably not random and that Meconopsis-Papaver species are prone to overcome a triploid block and establish stable lineages by dysploidy formation (2n = 21→2n = 20/22).

Autoploidy within Meconopsis

In addition to Meconopsis sect. Primulinae and M. sect. Grandes, sequences of M. sect. Aculeatae also fell into two different clades (Clade 1 and Clade 3, Fig. 3), suggesting that M. sect. Aculeatae might also have a hybrid history. The reconstructed polyploid evolutionary scenario suggests that Meconopsis sect. Aculeatae and M. sect. Meconopsis acquired chromosome levels of 2n = 56 independently, which is strongly supported by the aberrant chromosome number 2n = 28 in M. sect. Meconopsis and 2n = 14 in M. sect. Aculeatae (Fig. 2B). This conclusion contradicts a previous assumption, postulated without any available phylogenetic reference at that time, by Ratter (1968) that species in M. sect. Aculeatae and M. sect. Meconopsis were derived from a common ancestor with 2n = 56. However, sequences of M. sect. Meconopsis are only in Clade 2 with no strong divergence as shown in other Meconopsis sections, thus, it is possible that the high ploidy level in M. sect. Meconopsis resulted via autoploidy. For easy visualization, we simplified GAPDH network and displayed it in a reticulate network form (Fig. 4), which illustrates and summarizes our hypotheses of reticulate evolution of Meconopsis.

In conclusion, we postulate that there were three major pathways early in the evolution of Meconopsis: (1) formation of a triploid hybrid that managed to overcome sterility at the base of the clade that gave rise to M. sect. Primulae and M. sect. Grandes; (2) a hybrid origin of M. sect. Aculeatae with successive polyploidizations; (3) a autoploidization origin of M. sect. Meconopsis. Future research will be needed to test these hypotheses further, perhaps using next generation sequencing technique to generate more nuclear sequence or transcriptome data. Nonetheless, this is the first hypothesis of ancient hybridization and polyploidy underlying the evolution of Meconopsis.

Acknowledgements

We would like to thank the reviewers for their insightful suggestions and providing additional literature, which greatly helped to improve the clarity of result presentation and deepen our discussion.

Literature Cited

1.

Bretagnolle, F. A. and J. Thompson. 1995. Gametes with the somatic chromosome number: Mechanisms of their formation and role in the evolution of autopolypoid plants. New Phytol. 1: 1–22. Google Scholar

2.

Carolan, J. C., I. L. Hook, M. W. Chase, J. W. Kadereit, and T. R. Hodkinson. 2006. Phylogenetics of Papaver and related genera based on DNA sequences from its nuclear ribosomal DNA and plastid trnL intron and trnL-F intergenic spacers. Ann. Bot. 98: 141–155. Google Scholar

3.

Cobb, J. L. S. 1984. Meconopsis hybrids. Bull. Alpine Gard. Soc. Gr. Brit. 52: 63–73. Google Scholar

4.

Comai, L. 2005. The advantages and disadvantages of being polyploidy. Nat. Rev. Genet. 6: 836–846. Google Scholar

5.

Huelsenbeck, J. and F. Ronquist. 2001. Mrbayes: Bayesian inference of phylogenetic trees. Bioinformatics 17: 754–755. Google Scholar

6.

Kadereit, J. W. 1986. Papaver somniferum L. (Papaveraceae): a triploid hybrid? Bot. Jahrb. Syst. 106: 221–244. Google Scholar

7.

Kadereit, J. W. 1987. Experimental evidence on the affinities of Papaver somniferum (Papaveraceae). Pl. Syst. Evol. 156: 189–195. Google Scholar

8.

Kadereit, J. W. 1991. A note on the genomic consequences of regular bivalent formation and continued fertility in triploids. Pl. Syst. Evol. 175: 93–99. Google Scholar

9.

Kumar, S., S. M. Jeelani, S. Rani, R. C. Gupta, and S. Kumari. 2013. Cytology of five species of subfamily Papaveroideae from the western Himalayas. Protoplasma 250: 307–316. Google Scholar

10.

Lavania, U. and S. Srivastava. 1999. Quantitative delineation of karyotype variation in Papaver as a measure of phylogenetic differentiation and origin. Curr. Sci. 77: 429–435. Google Scholar

11.

Martin, D. P., D. Posada, K. A. Crandall, and C. Williamson. 2005. A modified bootscan algorithm for automated identification of recombinant sequences and recombination breakpoints. AIDS Res. Hum. Retroviruses 21: 98–102. Google Scholar

12.

Mayrose, I., M. S. Barker, and S. P. Otto. 2010. Probabilistic models of chromosome number evolution and the inference of polyploidy. Syst. Biol. 59: 132–144. Google Scholar

13.

McNaughton, I. 2014. The importance of polyploidy in Meconopsis with particular reference to the big perennial blue poppies. Sibbaldia 11: 159–173. Google Scholar

14.

Nessler, C. L. 1994. Sequence analysis of two new members of the major latex protein gene family supports the triploid-hybrid origin of the opium poppy. Gene 139: 207–209. Google Scholar

15.

Posada, D. 2008. Jmodeltest: Phylogenetic model averaging. Molec. Biol. Evol. 25: 1253–1256. Google Scholar

16.

Prain, D. 1915. Some additional species of Meconopsis. Bull. Misc. Inform. Kew 1915: 129–177. Google Scholar

17.

Ramsey, J. and D. W. Schemske. 1998. Pathways, mechanisms, and rates of polyploid formation in flowering plants. Ann. Rev. Ecol. Syst. 29: 467–501. Google Scholar

18.

Ratter, J. 1968. Cytological studies in Meconopsis. Notes Roy. Bot. Gard. Edinburgh, 28: 191–200. Google Scholar

19.

Rieseberg, L. H. and J. H. Willis. 2007. Plant speciation. Science 317: 910–914. Google Scholar

20.

Stamatakis, A. 2006. Raxml-vi-hpc: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–2690. Google Scholar

21.

Stamatakis, A., P. Hoover, and J. Rougemont. 2008. A rapid bootstrap algorithm for the raxml web servers. Syst. Biol. 57: 758–771. Google Scholar

22.

Taylor, G. 1934. An account of the genus Meconopsis. London: New flora and silva Limited. Google Scholar

23.

Treier, U., O. Broennimann, S. Normand, A. Guisan, U. Schaffner, T. Steinger, and H. Müller-Schärer. 2009. Shift in cytotype frequency and niche space in the invasive plant Centaurea maculosa. Ecology 90: 1366–1377. Google Scholar

24.

Xiao, W. 2013. Molecular systematics of Meconopsis Vig. (Papaveraceae): taxonomy, polyploidy evolution, and historical biogeography from a phylogenetic insight. Ph. D. Dissertation, The University of Texas at Austin. Google Scholar

25.

Ying, M., H. Xie, Z. Nie, Z. Gu, and Y. Yang. 2006. A karyomorphological study on four species of Meconopsis vig.(Papaveraceae) from the Hengduan Mountains, SW China. Caryologia 59: 1–6. Google Scholar

Appendices

Appendix 1. Voucher and sequence information (Accession number; species name; voucher (herbarium); (Collecting) COUNTRY: Subdivision; GenBank ID for matK, ndhF, trnL-trnF, rbcL. “-”, denotes a missing sequence).

X003; Meconopsis dhwojii G. Taylor; UK (cultivated); W. Xiao RICB9 (E); JX087915, JX087815, JX087755, JX087699. X004; Meconopsis wallichii Hook.; UK (cultivated); W. Xiao RICB10 (E); JX087895, JX087821, -, JX087711. X005; Meconopsis paniculata Prain; UK (cultivated); W. Xiao RICB5 (E); JX087868, JX087830, JX087743, JX087720. X006; Meconopsis superba King ex Prain; UK (cultivated); W. Xiao RICB7 (E); JX087858, JX087851, JX087735, JX087683. X007; Meconopsis simplicifolia (D. Don) Walp.; NEPAL: Bagmati; Egan 4 (private collection); JX087891, JX087803, JX087751, JX087700. X008; Meconopsis grandis Prain; UK (cultivated); W. Xiao RICB6 (E); JX087873, JX087832, -, JX087695. X009; Meconopsis betonicifolia Franch.; UK (cultivated); W. Xiao RICB2 (E); JX087871, JX087806, -, JX087716. X010; Meconopsis integrifolia (Maxim.) Franch.; CHINA: Yunnan; W. Xiao 080620 (TEX); JX087901, JX087804, -, JX087701. X011; Meconopsis horridula Hook.f. & Thomson; CHINA: Sichuan; Boufford 33724 (GH); JX087905, JX087812, JX087770, JX087712. X012; Meconopsis horridula Hook.f. & Thomson; CHINA: Yunnan; W. Xiao 080616 (TEX); JX087898, JX087826, -, JX087729. X015; Meconopsis punicea Maxim.; CHINA: Sichuan; Boufford 33684 (GH); JX087862, JX087849, -, JX087718. X016; Meconopsis quintuplinervia Regel; CHINA: Sichuan; W. Xiao RICB8 (E); JX087865, JX087831, -, JX087706. X018; Meconopsis lancifolia Franch. ex Prain; CHINA: Yunnan; W. Xiao 080621-1 (TEX); JX087857, JX087818, JX087750, JX087731. X019; Meconopsis henrici Bureau & Franch.; CHINA: Sichuan; W. Xiao 090722-1 (TEX); JX087913, JX087797, JX087739, JX087724. X020; Meconopsis speciosa Prain; CHINA: Yunnan; W. Xiao 090703-2 (TEX); JX087920, JX087829, JX087781, JX087682. X022; Meconopsis delavayi Franch. ex Prain; UK (cultivated); W. Xiao 090526 (TEX); JX087866, JX087816, JX087736, JX087688. X024; Cathcartia oliveriana (Franch. ex Prain) W. Xiao; CHINA: Shaanxi; J.Z. Xiao 1 (TEX); JX087907, JX087791, JX087765, -. X026; Meconopsis aculeata Royle; UK (cultivated); C5255 (E); JX087912, JX087820, -, JX087709. X027; Meconopsis bella Prain; NEPAL: Kone Khola; McBeath 1496 (E); JX087919, JX087823, -, JX087723. X028; Meconopsis torquata Prain; CHINA: Xizang; Ludlow 9904 (E); JX087875, -, JX087737, JX087696. X029; Meconopsis forrestii Prain; CHINA: Yunnan; Fang1154 (Xiang Ge Li La Alpine Garden); JX087853, JX087807, JX087734, -. X032; Meconopsis sp; CHINA: Sichuan; Boufford 33308 (GH); JX087903, JX087837, JX087749, JX087710. X034; Cathcartia chelidonifolia (Bureau & Franch.) W. Xiao; UK (cultivated); W. Xiao RICB4 (E); JX087897, JX087840, -, JX087690. X036; Meconopsis discigera Prain; BHUTAN: Upper Mo Chu District; Bowes Lyon15045 (E); JX087918, JX087824, JX087774, JX087686. X042; Meconopsis sinuata Prain; INDIA: Sikkim; ESK 683 (E); JX087890, JX087785, -, JX087725. X045; Meconopsis wumungensis K.M. Feng; CHINA: Yunnan; Liu 1990July (KUN); JX087922, -, -, JX087707. X046; Meconopsis wilsonii Grey-Wilson; CHINA: Sichuan; Boufford 32733 (GH); JX087924, JX087838, JX087740, JX087691. X047; Meconopsis primulina Prain; BHUTAN: Upper Mo Chu District; Sargent170 (E); JX087887, JX087843, -, JX087685. X052; Meconopsis concinna Prain; CHINA: Yunnan; Boufford 35133 (GH); JX087889, JX087841, JX087759, JX087721. X054; Meconopsis x cookei G. Taylor; CHINA: Qinghai; Long 696 (E); JX087869, JX087827, -, JX087726. X055; Cathcartia villosa Hook.f.; INDIA: Sikkim; ESK 205 (E); -, JX087847, -, JX087708. X069; Meconopsis autumnalis P.A. Egan; NEPAL: Bagmati; Egan 17 (private collection); JX087872, JX087822, JX087748, JX087714. X073; Meconopsis lyrata (H.A. Cummins & Prain) Fedde; BURMAR: N.E. upper Burma; Forrest 25047 (E); -, JX087800, -, -. X083; Meconopsis pseudovenusta G. Taylor; CHINA: Yunnan; W. Xiao 090705-2 (TEX); JX087894, JX087796, JX087741, -.

Appendix 2. Species name and Genbank ID for the downloaded trnL-trnF sequences (previously published by other authors).

Roemeria refracta DC.; DQ251150.1. Meconopsis latifolia Prain; AY328226.1. Papaver pavoninum C.A.Mey.; DQ251134.1. Papaver argemone L.; DQ251149.1. Papaver hybridum L.; DQ251152.1. Papaver apulum Ten.; DQ251151.1. Papaver heterophyllum Greene; DQ251146.1. Papaver californicum A.Gray; DQ251169.1. Papaver aculeatum Thunb.; DQ251168.1. Papaver spicatum Boiss. & Balansa; AY328244.1. Papaver cambricum L.; DQ251128.1. Papaver pilosum Sm.; DQ251172.1. Papaver armeniacum Lam.; DQ251148.1. Papaver orientale L.; DQ251143.1. Papaver bracteatum Lindl.; DQ251138.1. Papaver pseudo-orientale (Fedde) Medw.; DQ251147.1. Papaver rupifragum Boiss. & Reut.; DQ251165.1. Papaver atlanticum (Ball) Coss.; DQ251154.1. Papaver dubium L.; DQ251121.1. Papaver somniferum L.; DQ251132.1. Papaver macrostomum Boiss. & A.Huet; DQ251126.1. Papaver rhoeas L.; FJ626566.1. Papaver triniaefolium Boiss.; AM397153.1. Papaver commutatum Fisch., C.A.Mey. & Trautv.; DQ251164.1. Papaver glaucum Boiss. & Hausskn. ex Boiss.; DQ251159.1. Papaver alpinum L.; DQ251119.1. Papaver nudicaule L.; DQ251135.1. Papaver radicatum Rottb. ex DC.; DQ251113.1.

Appendix 3. GAPDH sequences information.

i1097-993X-17-1-5-t02.tif
Copyright 2014 by the Plant Resources Center, The University of Texas at Austin.
Wei Xiao and Beryl B. Simpson "The Role of Allotriploidy in the Evolution of Meconopsis (Papaveraceae): A Preliminary Study of Ancient Polyploid and Hybrid Speciation," Lundellia 17(1), 5-17, (1 December 2014). https://doi.org/10.25224/1097-993X-17.1.5
Published: 1 December 2014
KEYWORDS
ancestral chromosome number
Himalaya
hybridization
low-copy marker
polyploidization
speciation mechanism
Back to Top