Next Article in Journal
Survival and Genome Diversity of Vibrio parahaemolyticus Isolated from Edible Aquatic Animals
Next Article in Special Issue
Effects of Bamboo Forest Type and Density on the Growth of Bletilla striata and Root Endophytic Fungi
Previous Article in Journal
Identification and Characterization of UDP-Glycosyltransferase Genes in a Cerambycid Beetle, Pharsalia antennata Gahan, 1894 (Coleoptera: Cerambycidae)
Previous Article in Special Issue
Flora and Vegetation of Yunnan, Southwestern China: Diversity, Origin and Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evolution of Reproductive Traits and Implications for Adaptation and Diversification in the Yam Genus Dioscorea L.

1
Institute of Botany, Jiangsu Province and Chinese Academy of Sciences, Nanjing 210014, China
2
College of Horticulture, Academy for Advanced Interdisciplinary Studies, Nanjing Agricultural University, Nanjing 210014, China
*
Author to whom correspondence should be addressed.
Diversity 2022, 14(5), 349; https://doi.org/10.3390/d14050349
Submission received: 19 April 2022 / Revised: 27 April 2022 / Accepted: 27 April 2022 / Published: 29 April 2022
(This article belongs to the Special Issue Ecology, Evolution and Diversity of Plants)

Abstract

:
Dioscorea is a pantropical monocotyledonous genus encompassing several well-known tuber crops and medicinal plants. It possesses remarkable morphological diversity, especially in reproductive characteristics, which are suggested to play important roles in species adaptation and diversification. Yet there have been few studies that consider the evolutionary pattern followed by these characters in this genus. In this study, the phylogenetic relationships among Chinese yams were reconstructed from five chloroplast and two mitochondrial DNA sequences. The evolutionary histories of bulbil possession, inflorescence architecture, the color of the male flowers and the degree of male flower opening were reconstructed. The results suggested that yam bulbils evolved after the divergence between D. sect. Testudinaria and other species of Dioscorea except for in D. sect. Stenophora and D. sect. Apodostemon. The evolutionary trend in the degree of male flower opening ranged from fully open to nearly closed. Male flowers with dark colors and panicles were shown to be derived in Dioscorea. These characteristics were found to be closely associated with the reproductive patterns and pollinating mechanisms of the Dioscorea species. The findings also shed light on the systematic relationships within this genus.

1. Introduction

Dioscorea L. is the largest genus of the Dioscoreaceae, comprising about 630 species [1], mainly distributed in tropical and subtropical areas, sometimes expanding to temperate regions [2]. Species of this genus are mostly dioecious vines with underground storage organs [3]. The edible starchy tuber makes it the third most important tropical tuberous crop globally [4], and the metabolite-rich rhizomes of some species are used as a source of pharmaceutical compounds [5]. Apart from its economic importance, Dioscorea is one of the most critical taxa in monocot systematics since it is laid near the basal position in the phylogenetic tree of monocotyledonous plants and has a number of similarities with dicots [6,7]. However, there are challenges in the taxonomic and systematic investigation of Dioscorea due to its great morphological diversity, dioecy and small flowers [2,8]. There are about 1600 taxonomic names attributed to Dioscorea, but most of them are considered synonyms [1], indicating the controversy on the circumscription of species boundaries. Moreover, the classification system of Dioscorea differs greatly among authors, and many of the proposed infrageneric taxa do not quite represent natural lineages [2]. Based on morphological and anatomic characteristics, Dioscorea species have been assigned to between 24 and 58 sections in traditional taxonomic studies [3,9,10,11]. Recent phylogenetic analyses based on molecular data sets, including cpDNA regions and nuclear genes, have led to the recognition of 10–11 main clades in this genus [2,12,13,14,15]. The results greatly simplified previous classification systems and gave rise to a fundamental phylogenetic framework for understanding the infrageneric relationships and evolutionary history of Dioscorea. Despite extensive study of phylogeny and biogeography, research focusing on character evolution and its relation to species diversification and ecological success are extremely limited.
During angiosperm evolution, advances in vegetative and reproductive organs generated remarkable morphological diversity and were pivotal to niche adaptation and speciation [16]. Recent phylogenetic research has indicated some lineages of Dioscorea, such as the Madagascar group and D. sect. Enantiophyllum, have very short phylogenetic branches, resulting from radiation events [2,12,13,17]. It has been proposed that fast evolution over a short period may reflect responses to tremendous climate change or the rise of innovative characters [18]. Dioscorea species exhibit extreme morphological variations; for example, underground storage organs include branching rhizomes, perennial or annual tubers, cylindrical or globose tubers, right or left stem twining and seed wings extending from the apex, the base or around the whole seed. Previous studies revealed the distribution patterns of some of these traits, including tuber morphology, stem twining direction, seed wing shape [2], stem anatomy [8], leaf venation [19], pollen characters and chromosome numbers [15]. However, these studies mainly focused on identifying synapomorphies for infrageneric lineages rather than elucidating the evolutionary history of traits and their links with the wide range of environmental conditions in which the plants lived.
Burkill [11] suggested that adaptation to limiting rainfall was the most important factor underpinning species diversification in Dioscorea. The restricted distribution and fewer species in D. sect. Stenophora, compared with great diversity and pantropical distribution of other clades of Dioscorea, led Wilkin et al. [2] to propose that tubers play an important role in the origins of diversity in this genus. Investigations focusing on African Dioscorea species showed that the shift from forest to open grassland was associated with changes in tuber size and orientation that protect the plants from fires. Their stem habit shifted from twining to erect due to the lack of supporting vegetation, and seed wing morphology adapted to release at low height, requiring higher wind speeds for efficient dispersal [20]. In addition, changes in flower and fruit morphology were suggested to play key roles in the exposure to radiation of the Madagascan clade [2]. These morphological traits can greatly influence the pollination process and seed dispersal, further altering the fitness of different species in given conditions. However, the evolutionary pattern of reproductive characters in Dioscorea, especially floral traits, remain poorly understood.
The Himalayan-Hengduan Mountains are purported to be the center of origin and diversification for the yam genus, with a large proportion of endemic species [21]. There are 52 species, 1 subspecies and 9 varieties of Dioscorea in China, which show enormous variability in traits related to sexual and vegetative reproduction [22]. For instance, some species produce bulbils at the leaf axils, while others do not; the inflorescence may be spikes, racemes or panicles; the color of male flowers varies between white, yellow, green to orange, purple and so on; and the perianths of male flowers are completely open, half-closed or fully closed during blooming in different species (shown in Figure 1). These features are widely used to divide plants into infrageneric groups and define species. This group of Dioscorea species is, therefore, an excellent candidate for exploring character evolution and implications for species diversification and ecological adaptation. In this study, we reconstructed the phylogenetic relationships between 48 Dioscorea species using 5 chloroplast and 2 mitochondrial DNA markers. Based on the phylogenetic framework produced, we explored the evolutionary patterns of four reproductive characters to clarify the driving force for the diversification and adaptive evolution of this important angiosperm lineage.

2. Materials and Methods

2.1. Sampling

A total of 48 Dioscorea taxa were sampled in this study, covering all sections distributed in China, except the monotypic section D. sect. Stenocorea. The samples consisted of 43 species, 1 subspecies and 4 varieties. Among them, 47 taxa were collected in the field and the vouchers were deposited in the Herbarium of the Institute of Botany, Jiangsu Province and the Chinese Academy of Sciences, China (NAS). The sequences of the remaining species, D. wallichii J. D. Hooker, were downloaded from GenBank. We selected Tacca chantrieri Andre (Dioscoreaceae), a representative of the genus closely related to Dioscorea according to the recently built phylogeny of the Dioscoreaceae [14], as the outgroup in our analyses. The geographical origin, voucher specimen information and GenBank accession numbers of all samples are listed in Table 1.

2.2. DNA Extraction, Amplification and Sequencing

The total genomic DNA was extracted from fresh or silica-gel dried leaves following a modified CTAB method [23] and stored at −20 °C before amplification. Five chloroplast markers, matK, rbcL, trnL-F, psbA-trnH and rpl36-rps8, plus two mitochondrial markers, nad1 and rps3, were chosen, taking evolutionary rate and ease of sequencing into account. The primers used for matK and rbcL amplification followed Gao et al.’s recommendations [24]. The amplification of trnL-F used primers described by Taberlet et al. [25]. The primers for psbA-trnH and rpl36-rps8 were newly designed based on homologous sequences of D. elephantipes (L’Hér.) Engl. on GenBank. The primers for the nad1 gene were designed based on the sequence of Oryza sativa Linn. The rps3 gene was amplified according to Laroche and Bousquet [26]. The sequences of all primers used in this study are detailed in Table 2.
The PCR reaction mixture contained 40 ng of genomic DNA template, 2.5 mmol/L MgCl2, 1 × Mg-free DNA polymerase buffer, 0.12 mmol/L dNTPs, 0.3 mmol/L of each primer, 1 U Taq DNA polymerase and water added accordingly to a final volume of 50 μL. The PCR program for matK, rbcL, trnL-F, nad1 and rps3 was as follows: a 3 min premelt at 94 °C, followed by 35 cycles of 45 s denaturation at 94 °C, 30 s annealing at 58 °C and a 1.5 min extension at 72 °C, plus a final extension of 5 min at 72 °C. For psbA-trnH and rpl36-rps8, the PCR reaction included an initial denaturation at 94 °C for 3 min, followed by 35 cycles of denaturation at 94 °C for 30 s, annealing at 54 °C for 30 s, extension at 72 °C for 80 s and a final extension at 72 °C for 5 min. The PCR products were examined electrophoretically using 0.8–1.2% agarose gels and purified using a TIAN gel Midi Purification Kit (TIANGEN Biotech, Beijing, China). The purified products were sequenced using the same primer pairs with PCR. All of the sequences obtained were subjected to a BLAST (https://blast.ncbi.nlm.nih.gov/Blast.cgi (accessed on 15 April 2020)) search to detect contamination and nonspecific amplification. After confirmation, the newly generated sequences were submitted to GenBank.

2.3. Phylogenetic Analyses

The assembled sequences were aligned using MAFFT v.7 [27] with default settings and then adjusted manually for accuracy in Geneious R9 9.1.8 (https://www.geneious.com (accessed on 4 July 2020)). Character gaps were treated as missing data. Phylogenetic relationships were reconstructed using maximum likelihood (ML), maximum parsimony (MP) and Bayesian inference (BI) approaches. ML analysis was firstly performed to build a single-gene tree in online CIPRES Science Gateway v.3.3 (http://www.phylo.org/ (accessed on 18 Octobor 2020)) [28], using RAxML-HPC v8.2.10 [29,30] under the GTR + G DNA substitution model [31]. One hundred rapid bootstrap replicates were generated for each single gene matrix [32]. Since there was no strong conflict among individual markers, we conducted further ML analysis using a concatenated matrix of all seven loci under the GTR + G model as recommended by jModelTest v2.1 [33]. Bootstrap analyses were used to evaluate the support for each clade with 1000 bootstrap replicates. MP analysis was conducted in PAUP* version 4.0b10 [34]. All characters were equally weighted. Trees were inferred using the heuristic search option with tree bisection probabilities (TBR) swapping and 1000 replicates of random addition. Ten trees were held in each step during stepwise addition. The maximum number of trees was set to 10,000, and all parsimonious trees were saved. The tree length (TL), consistency index (CI), retention index (RI) and rescaled consistency index (RC) were calculated for each maximum parsimony tree. The bootstrap results were summarized in a 50% majority-rule consensus cladogram. BI analysis was implemented using MrBayes version 3.2.6. [35]. A Markov Chain Monte Carlo (MCMC) analysis was run with two independent chains with a random starting tree for 100 million generations, sampling one tree every 1000 generations. The first 25% generations were discarded as burn-in, and the remaining trees were used to construct a consensus tree with a 50% majority rule and obtain the posterior probabilities.

2.4. Character Coding and Ancestral State Reconstruction

Four characters closely related to vegetative or sexual reproduction were selected for ancestral character state reconstruction, namely bulbil formation, inflorescence structure, floral color and opening degree of the perianth. Morphological data were obtained from direct observations and the literature. The evolutionary history of each of the four characters was traced over the Bayesian 50% majority-rule tree using MP approaches available in Mesquite 3.5.1 [36]. The character states were treated as unordered and equally weighted. Morphological characters and their states were coded as follows: a. bulbil absence (0), presence (1); b. inflorescence umbel (0), spike (1), raceme (2) or panicle (3); c. floral color dark (e.g., violet, purplish red, orange) (0) or light (e.g., white, pale yellow, green) (1); d. perianth opening degree open (0), half-closed (1) or closed (2). For species in which the floral color changed during blooming, we used the color at the mature stage as the character state. The resulting codes for the sampled species are summarized in Table 3.

3. Results

3.1. Phylogenetic Analyses

According to our analyses, the chloroplast segment trnL-F exhibited the highest variability among the seven molecular markers, followed by matK and nad1 (see Table 4). The final concatenated matrix of all seven markers consisted of 7218 base pairs containing 1157 variable sites and 614 parsimony-informative sites. The molecular reconstruction based on the combined matrix gave rise to a highly resolved phylogenetic tree with moderate to strong support. There was no significant incongruence in topology or support values among trees generated from different methods. Thus, only the ML tree is presented in Figure 2.
All sampled Dioscorea species clustered in a monophyletic group consisting of two major clades with strong support. In accordance with previous studies, D. sect. Stenophora was the first diverging clade, sister to the other clade that included all the remaining species (BS = 100). The species of D. sect. Botryosicyos was grouped into a monophyletic clade (BS = 100). This clade was further divided into two subclades, one consisting of D. pentaphylla L., D. esquirolii Prain et Burkill and D. menglaensis H. Li and the other including D. melanophyma Prain et Burkill, D. kamoonensis Kunth and D. delavayi Franch. The only species of D. sect. Lasiophyton, D. hispida Dennst, was sister to D. sect. Botryosicyos (BS = 93). The position of D. tentaculigera Prain et Burkill remained unresolved, while other species of D. sect. Shannicorea clustered together in a well-supported clade (BS = 100). The two samples of D. sect. Combilium grouped together, and they were revealed as sisters to the Shannicorea clade with moderate support (BS = 83). All accessions of D. sect. Enantiophyllum formed a well-supported monophyletic clade (BS = 100). The species of D. sect. Opsophyton (D. bulbifera) was positioned as a sister to D. sect. Enantiophyllum with moderate support (BS = 71).
Some taxonomic treatments at the species level were supported by the phylogenetic tree. Dioscorea collettii Hook. f. and D. collettii var. hypoglauca (Palibin) C. T. Ting grouped together, while D. esculenta (Lour.) Burkill and D. esculenta var. spinosa (Roxburgh ex Prain & Burkill) R. Knuth were resolved as the closest relatives to each other. However, some varieties or subspecies of one species did not form a monophyletic group. For instance, D. nipponica Makino and D. nipponica subsp. rosthornii (Prain & Burkill) C. T. Ting, and D. subcalva Prain et Burkill and D. subcalva var. submollis (R. Knuth) C. T. Ting & P. P. Ling, did not group together.

3.2. Ancestral Character State Analyses

The ancestral state of four reproductive features of the Dioscorea species was reconstructed based on MP methods. As for character “a” the absence or presence of bulbils at the axil), state “0” (absent) was suggested to be the plesiomorphic state for Dioscorea (Figure 3). No species of the outgroup Tacca and D. sect. Stenophora, the earliest diverging group in Dioscorea, produces bulbils. The trait distribution in the phylogenetic tree suggests that bulbils appeared after the divergence of D. sect. Stenophora, but were subsequently lost several times independently, in D. sect. Combilium, D. sect. Lasiophyton and several species in D. sect. Enantiophyllum (D. wallichii Hook. f., D. aspersa Prain et Burkill). Bulbils persisted in most species in D. sect. Botryosicyos, D. sect. Opsophyton and D. sect. Enantiophyllum.
The ancestral state of the inflorescence architecture was inferred to be a spike (Figure 4). Racemes and panicles were inferred to be derived states, with the former occurring sporadically in different clades and the latter occurring mainly in D. sect. Botryosicyos and D. sect. Enantiophyllum. Spike occurred as a homoplasy in D. sect. Shannicorea, while other clades did not show a consistent type of male inflorescence.
The light color of the male flowers was uncovered as the ancestral state of Dioscorea (Figure 5). The dark color evolved independently several times in D. sect. Stenophora and D. sect. Opsophyton. Most species with dark colored male flowers occurred in D. sect. Stenophora. There were also some species in D. sect. Shannicorea and D. sect. Enantiophyllum that possessed male flowers that were purple in color or had a brown stripe on them. This distribution probably reflects multiple convergence events in response to similar pollination strategies.
The male flowers of Dioscorea ancestors were suggested to open fully, and this state was retained in D. sect. Stenophora and D. sect. Shannicorea (Figure 6). The perianth of male flowers was closed to some degree in the ancestor of the clade encompassing D. sect. Enantiophyllum, D. sect. Opsophyton, D. sect. Botryosicyos, D. sect. Lasiophyton and D. sect. Combilium. Male flowers of D. sect. Botryosicyos and D. sect. Enantiophyllum were generally closed.

4. Discussion

4.1. Evolution of Organs for Vegetative Propagation and Diversity in the Reproductive Strategy

Vegetative propagation is a crucial mechanism that ensures plants maintain and spread their population, allowing plants to survive unfavorable conditions for sexual reproduction [37]. Dioscorea species possess multiple organs involved in vegetative propagation, including rhizomes, horizontal or vertical tubers and aerial bulbils, as a result of adaptation to various environmental conditions. In the early diverged D. sect. Stenophora, the underground parts consisted of perennial branched horizontal rhizomes. Perennial to annual starchy tubers took the place of rhizomes in subsequently evolved lineages to undertake the function of asexual reproduction. Rhizomatous species are supposed to be more adaptive to long growing seasons in shaded habitats [38], whereas tubers can evade destruction by animals and allow adaptation to warm and humid climates [39]. The possession of tubers was suggested to play an important role in species diversification of Dioscorea [2]. However, tubers are mainly storage organs for plants to live through winter and germinate in for the next growing season; the reproductive efficiency and dispersal ability of tubers are limited.
The emergence of aerial bulbils in some Dioscorea species greatly changed the propagation patterns of these species. The bulbil of Dioscorea is a minor storage organ compared to the underground tuber, arising from the axils of leaves or inflorescences and varying in form and size in different species [40]. On the basis of morphological and anatomical studies, bulbils are interpreted as modified axillary branches and miniature reproductions of the rhizome [11,41]. Functioning as a means of vegetative propagation, bulbils are released from the plants after maturation and then dispersed by gravity or by water in streams, sometimes over long distances, like seeds. Bulbils can easily germinate and grow independently, giving rise to large numbers of progeny. Because of these advantageous characteristics, species that produce bulbils are usually more vigorous and more competitive than species that reproduce only by seeds, as observed in Allium (Alliaceae) [42] and Butomus (Butomaceae) [43].
As mentioned above, the vegetative propagation strategies vary among the different lineages of Dioscorea, and the production of bulbils is a derived character in most clades of this genus. This may have greatly influenced the expansion of the Dioscorea species. In Poa alpina L. (Poaceae), bulbil-producing plants are better adapted to higher elevations compared with seed-producing plants. Therefore, this species is able to occupy a range of ecological niches by means of different reproductive modes [44]. Bulbils in Fritillaria (Liliaceae) are suspected to be an adaptation to underground dispersal [38]. In Dioscorea, the rhizomatous taxa are mainly distributed in South and East Asia, and many of them are endemic species only found in restricted areas, while the species that produce bulbils occupy a wide range of habitats worldwide [2]. D. bulbifera, the only wild species exhibiting worldwide distribution in Dioscorea, produces huge numbers of bulbils compared to other species. Bulbils of D. sansibarensis Pax are buoyant and spread through water flow, achieving wide distribution in African valleys [11]. Here, it is postulated that the production of bulbils improved the adaptive ability of Dioscorea species and promoted their expansion.

4.2. Evolution of Floral Characters and Their Relation to Sexual Propagation

Flower evolution is strictly linked to pollination strategy [45,46]. Dioscorea is predominantly dioecious and allogamous; thus sexual propagation in this genus relies on the outcrossing process. Moreover, the sticky nature of Dioscorea pollen grains prevents transport by wind, and, consequently, the flowers of Dioscorea species are pollinated mainly by insects [47]. The diverse flower exhibition characteristics of this genus were most likely shaped by pollinator-mediated selection.
According to our analyses, spikes are reconstructed as the plesiomorphy of Dioscorea, while panicles appear mainly in D. sect. Enantiophyllum as a derived inflorescence type. This contrasts with the opinion presented by Burkill [11], who suggested that spikes and racemes evolved from panicles. The distribution patterns of inflorescence types among species do not mirror phylogenetic relationships, with closely related lineages displaying different types, except D. sect. Shannicorea. Such a pronounced polymorphism might have been driven by pollinators. Inflorescence architecture is related to the mode and efficiency of pollination. For instance, changes in inflorescence architecture were associated with the transition from biotic (insect) to abiotic (wind) pollination in Schiedea salicaria Hbd. (Caryophyllaceae) [48], and affected pollinator behavior and mating success in Spiranthes sinensis (Pers.) Ames (Orchidaceae) [49]. Panicles are made of multiple spikes or racemes to enlarge the volume of the flower and enhance the attraction of pollinators. They possess two to three times the number of flowers and amount of pollen as spikes or racemes of the same length. The function of panicles in insect attraction compared with other types of inflorescences is worthy of investigation in Dioscorea.
In our results, most species of Dioscorea possess male flowers of light color, and dark color originated several times independently in different lineages. It is well known that floral color is among the most important visual signals in pollinator attraction. For example, hawkmoth and hummingbird pollinators prefer yellow and red morphs of Mimulus aurantiacus Curtis, respectively [50]. The reflectance spectrum of the dark perianth generated by UV-light is recognizable to insects [51]. Dark color and light color perianth could be visually discriminated by insects and thus will affect the visiting choice of different pollinators [52]. An alternative explanation for floral color polymorphisms among closely related species is that color divergence evolves in response to interspecific competition for pollinators as a means to decrease interspecific pollinator movements [53]. Indeed, the color of perianth in Dioscorea includes white, yellow, green, orange, red and purple, sometimes changing from one to another during flower development. How these variations of floral colors influence the category and behavior of the pollinators will be an interesting subject to investigate.
The evolutionary trend of the perianth opening degree in Dioscorea is from wholly opened to nearly closed. This character may also influence the category of pollinators. As discovered in the study of Merianieae (Melastomataceae), corollas of most buzz-bee syndrome species are widely open, forming bowl-shaped flowers, whereas they are more closed and form urceolate to pseudo-campanulate flowers in vertebrate-pollinated species [54]. The same was also observed in Erica (Ericaceae) [55] and Ruellia (Acanthaceae) [56]. Zhao et al. [57] reported that D. nipponica subsp. rosthornii is pollinated by halictids. In D. sect. Enantiophyllum species, male panicles were found usually to be made up of closed flowers that restrict visitations only to small insects. The pollinators of species such as D. rotundata Poir., D. japonica Thunb. and D. polystachya Turcz. were purported to be thrips [47,58,59,60]. Therefore, the evolution of male flower opening degree in Dioscorea may be a result of pollinator specialization.
Actually, as reported in previous studies, the pollinators of Dioscorea species include Coleoptera, Diptera, Hymenoptera, Hemiptera, Thripidae, Thysanoptera, Halictus and Andrena. [47,60]. The diversification of floral morphological characters in this genus is thought to be promoted by pollination-associated adaptations. Intensive investigation of the pollination strategies in other lineages is needed to better understand the adaptive evolution of reproduction in Dioscorea.

4.3. Infrageneric Relationships and Taxonomic States of Several Species

The phylogenetic framework reconstructed in this study was consistent with those obtained in previous studies [2,12,14,15]. The relationships indicated by molecular information were generally consistent with infrageneric division by traditional systematics.
The monophyly of Dioscorea sect. Stenophora has been certified in a number of molecular studies using various datasets. This is also supported by morphological evidence since species of D. sect. Stenophora show underground organ type, chromosome number and pollen type that set them apart from other lineages of Dioscorea. As indicated in this study, no species of D. sect. Stenophora generated bulbils, and they all had male flowers with fully open perianths, thus giving further support to the isolated position of this section within the Dioscorea genus.
Most species of Dioscorea sect. Lasiophyton have been transferred into D. sect. Botryosicyos [22]. However, there are clear morphological differences between these two taxa. Species of D. sect. Lasiophyton have compound leaves with three palmate leaflets that are also palmately veined, and all six stamens of the male flower are fertile; in contrast, in D. sect. Botryosicyos, leaves usually have more than three leaflets, are pinnately veined, and have three stamens alternating with three staminodes. In this study, the only Chinese species of D. sect. Lasiophyton, D. hispida, was resolved as the sister lineage to the Botryosicyos clade, supporting the exclusion of this species from D. sect. Botryosicyos by Ding and Gilbert [22].
Dioscorea tentaculigera was assigned to D. sect. Shannicorea by Burkill [11]. Its male flowers are light in color and fully open. However, D. tentaculigera produces bulbils, and its male flowers are sessile, which is an obvious departure from other species in D. sect. Shannicorea. The phylogenetic position of D. tentaculigera was not resolved in our analyses, as has been the case in previous studies (e.g., [2,12,13,14]). Maurin et al. [20] placed this species outside of the Enantiophyllum clade based on six cpDNAs; Soto Gomez et al. [61] positioned it as a sister to the Mediterranean clade based on 260 nuclear genes; and Noda et al. [15] suggested that it should be treated as a distinct section according to four cpDNAs. More evidence from molecular and morphological data is needed to draw a conclusion about the correct classification of this species.
Accessions of Dioscorea sect. Enantiophyllum were found to cluster into a highly supported monophyletic clade. There are several synapomorphies of this group: (i) stem twining to the right, (ii) leaves generally opposite, (iii) seeds winged all round, (iv) bulbils normally present, (v) male inflorescence panicles and (vi) perianth white to yellow and nearly closed when blooming. The limits of this section were found to be reasonable for a natural lineage [2,12,14,15,17].
The phylogenetic analyses also shed light on the taxonomy of certain species. Dioscorea nipponica subsp. rosthornii was separated from D. nipponica subsp. nipponica by a cork layer of rhizomes and male flowers that were pedicellate [62], as well as a difference in chromosome number [24,39]. However, they have not been resolved as sister groups in any molecular phylogenetic trees generated to date [12,14,24]. A similar situation was observed with D. subcalva var. submolis, which showed differences from D. subcalva var. subcalva that included sparse leaf blades and longer infructescences. We noticed that these characters varied in a continuous range between individuals among populations. The relationships between these two varieties and other species of D. sect. Shannicorea remain unresolved by molecular analyses [15,17]. Further investigation with sufficient samples may help us understand the internal relationships within this section and the circumscription of the species.

5. Conclusions

This study provides the first phylogenetic analyses focusing on the evolution of four main reproductive traits in Dioscorea. The development of bulbils in late-diverged lineages of Dioscorea has diversified the mode of vegetative propagation, which is supposed to have greatly improved reproductive efficiency and heightened plants’ ability to occupy new habitats. Spikes, male flowers in light color and wholly open perianth are reconstructed as plesiomorphic in Dioscorea, whereas panicles, dark color flowers and nearly closed perianth are suggested derived states. The extraordinary variation of floral characters in this genus appears to be a consequence of adaptive evolution driven by pollinators. A broader sampling of Dioscorea, especially new-world species, together with their character information and extensive pollination studies among species with different floral, types are necessary to achieve a better understanding of the diversification mechanism of this genus.
On the other hand, this work provides a suitable phylogenetic framework for the revision of infrageneric classification and species delimitation in Dioscorea. Some of the morphological characters show consistency in a certain clade. Species in D. sect. Stenophora do not produce bulbils, and their male flowers are always wholly open; D. sect. Shannicorea is characterized by no bulbils, spikes, light color and wholly open flowers; D. sect. Enantiophyllum is distinguished by producing bulbils, with flowers in light color and nearly closed. According to our results, bulbils and perianth opening degree are informative taxonomic features at the section level in Dioscorea. The monophyletic status of several sections proposed in classical taxonomy, such as D. sect. Stenophora and D. sect. Enantiophyllum, were supported by morphological synapomorphies besides molecular data. Morphological characters analyzed in the present work can also explain the unexpected position of D. tentaculigera in phylogenetic trees. In addition, our results suggest the necessity of reconsideration of some infraspecific taxa as well.

Author Contributions

Methodology, M.C. and J.-Y.X.; software, J.-Y.X.; validation, X.S., Y.H. and Y.Z.; formal analysis, M.C. and X.S.; investigation, M.C.; resources, Y.Z. and Y.H.; data curation, M.C.; writing—original draft preparation, M.C.; writing—review and editing, X.S. and Y.H.; supervision, Y.H.; project administration, Y.H.; funding acquisition, M.C. and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Youth Foundation of Jiangsu Province to M.C. (Grant No. BK20180316) and the Independent Research Project of Jiangsu Provincial Public Welfare Research Institute to X.Q.S. (Grant No. BM2018021-2).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data generated in this study have been uploaded to NCBI (https://www.ncbi.nlm.nih.gov/) and GenBank accession numbers are listed in Table 1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Govaerts, R.; Wilkin, P.; Raz, L.; Téllez-Valdés, O. World checklist of Dioscoreaceae. Facilitated by the Royal Botanic Gardens, Kew. 2018. Available online: https://wcsp.science.kew.org/ (accessed on 25 July 2021).
  2. Wilkin, P.; Schols, P.; Chase, M.W.; Chayamarit, K.; Furness, C.A.; Huysmans, S.; Rakotonasolo, F.; Smets, E.; Thapyai, C. A plastid gene phylogeny of the yam genus, Dioscorea: Roots, fruits and Madagascar. Syst. Bot. 2005, 30, 736–749. [Google Scholar] [CrossRef] [Green Version]
  3. Huber, H. Dioscoreaceae. Flowering plants. Monocotyledons. Lilianae (except Orchidaceae). In The Families and Genera of Vascular Plants; Kubitzk, K., Ed.; Springer: Berlin/Heidelberg, Germany, 1998; Volume 3, pp. 216–235. [Google Scholar]
  4. Asiedu, R.; Sartie, A. Crops that feed the world. Yams. Food Secur. 2010, 2, 305–315. [Google Scholar] [CrossRef]
  5. Price, E.J.; Wilkin, P.; Sarasan, V.; Fraser, P.D. Metabolite profiling of Dioscorea (yam) species reveals underutilized biodiversity and renewable sources for high-value compounds. Sci. Rep. 2016, 6, 29136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Dahlgren, G. An undated angiosperm classification. Bot. J. Linn. Soc. 1989, 100, 197–203. [Google Scholar] [CrossRef]
  7. Mignouna, H.; Abang, M.; Asiedu, R.; Geeta, R. True yams (Dioscorea): A biological and evolutionary link between eudicots and grasses. Cold Spring Harb. Protoc. 2009, 11, pdb.emo136. [Google Scholar] [CrossRef]
  8. Sheikh, N.; Kumar, Y. Foliar epidermal, stem and petiole anatomy of Meghalayan Dioscorea L. (Dioscoreaceae) and its systematic implication. Bangladesh J. Plant Taxon 2017, 24, 53–63. [Google Scholar] [CrossRef] [Green Version]
  9. Uline, E.B. Eine monographie de Dioscoreaceen. Bot. Jahrb. 1898, 25, 126–165. [Google Scholar]
  10. Knuth, R. Dioscoreaceae. In Das Pflanzenreich, 87 (IV 43); Engler, H.G.A., Ed.; Verlag von Wilhelm Engelmann (Druck von Breitkopf & Härtel in Leipzig): Leipzig, Germany, 1924; pp. 1–387. [Google Scholar]
  11. Burkill, I.H. The organography and the evolution of Dioscoreaceae, the family of the yams. Bot. J. Linn. Soc. 1960, 56, 319–412. [Google Scholar] [CrossRef]
  12. Viruel, J.; Segarra-Moragues, J.G.; Raz, L.; Forest, F.; Wilkin, P.; Sanmartín, I.; Catalán, P. Late cretaceous-early eocene origin of yams (Dioscorea, Dioscoreaceae) in the Laurasian Palaearctic and their subsequent oligocenemiocene diversification. J. Biogeogr. 2016, 43, 750–762. [Google Scholar] [CrossRef]
  13. Viruel, J.; Forest, F.; Paun, O.; Chase, M.W.; Devey, D.; Couto, R.S.; Gabriel, J.; Segarra-Moragues Catalán, P.; Wilkin, P. A nuclear Xdh phylogenetic analysis of yams (Dioscorea: Dioscoreaceae) congruent with plastid trees reveals a new Neotropical lineage. Bot. J. Linn. Soc. 2018, 187, 232–246. [Google Scholar] [CrossRef]
  14. Couto, R.S.; Martins, A.C.; Bolson, M.; Lopez, R.C.; Smidt, E.C.; Braga, J.M.A. Time calibrated tree of Dioscorea (Dioscoreaceae) indicates four origins of yams in the Neotropics since the Eocene. Bot. J. Lin. Soc. 2018, 188, 144–160. [Google Scholar] [CrossRef]
  15. Noda, H.; Yamashita, J.; Fuse, S.; Pooma, R.; Poopath, M.; Tobe, H.; Tamura, M.N. A large-scale phylogenetic analysis of Dioscorea (Dioscoreaceae), with reference to character evolution and subgeneric recognition. Acta Phytotax. Geobot. 2020, 71, 103–128. [Google Scholar]
  16. Endress, P.K. Diversity and Evolution Biology of Tropical Flowers; Cambridge University Press: Cambridge, UK, 1994; p. 511. [Google Scholar]
  17. Hsu, K.M.; Tsai, J.L.; Chen, M.Y.; Ku, H.M.; Liu, S.C. Molecular phylogeny of Dioscorea (Dioscoreaceae) in East and Southeast Asia. Blumea J. Plant Taxon. Plant Geogr. 2013, 58, 21–27. [Google Scholar] [CrossRef] [Green Version]
  18. Biffin, E.; Brodribb, T.J.; Hill, R.S.; Thomas, P.; Lowe, A.J. Leaf evolution in Southern Hemisphere conifers tracks the angiosperm ecological radiation. Proc. Biol Sci. 2012, 279, 341–348. [Google Scholar] [CrossRef]
  19. Sun, X.Q.; Xue, J.Y.; Lei, Z.; Li, M.M.; Zhang, Y.M.; Zhou, G.C.; Hang, Y.Y. Taxonomic and phylogenetic significance of leaf venation characteristics in Dioscorea plants. Arch. Biol. Sci. 2018, 70, 397–407. [Google Scholar] [CrossRef]
  20. Maurin, O.; Muasya, A.M.; Catalán, P.; Shongwe, E.Z.; Viruel, J.; Wilkin, P.; Van der Bank, M. Diversification into novel habitats in the Africa clade of Dioscorea (Dioscoreaceae): Erect habit and elephant’s foot tubers. BMC Evol. Biol. 2016, 16, 238. [Google Scholar] [CrossRef] [PubMed]
  21. Wan, J.R.; Ding, Z.Z.; Qin, H.Z. A phytogeographical study on the family Dioscoreaceae. Acta Bot. Boreale-Occident. Sin. 1994, 14, 128–135. [Google Scholar]
  22. Ding, Z.Z.; Gilbert, M.G. Dioscoreaceae. In Flora of China; Wu, Z.Y., Raven, P.H., Eds.; Science Press: Beijing, China; Botanical Garden Press: St. Louis, MI, USA, 2000; Volume 24, p. 284. [Google Scholar]
  23. Doyle, J.J.; Doyle, J.L. A rapid DNA isolation procedure for small amounts of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  24. Gao, X.; Zhu, Y.P.; Wu, B.C.; Zhao, Y.M.; Chen, J.Q.; Hang, Y.Y. Phylogeny of Dioscorea sect. Stenophora based on chloroplast matK, rbcL and trnL-F sequences. J. Syst. Evol. 2008, 46, 315–321. [Google Scholar]
  25. Taberlet, P.T.; Gielly, L.; Patou, G.; Bouvet, J. Universal primers for amplification of three non-coding regions of chloroplast DNA. Plant Mol. Biol. 1991, 17, 1105–1109. [Google Scholar] [CrossRef]
  26. Laroche, J.; Bousquet, J. Evolution of the mitochondrial rps3 intron in perennial and annual angiosperms and homology to nad5 intron 1. Mol. Biol. Evol. 1999, 16, 441–452. [Google Scholar] [CrossRef] [Green Version]
  27. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7, improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [Green Version]
  28. Miller, M.A.; Pfeiffer, W.; Schwartz, T. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In Proceedings of the Gateway Computing Environments Workshop (GCE), New Orleans, LA, USA, 14 November 2010; pp. 1–8. [Google Scholar]
  29. Stamatakis, A. RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 2006, 22, 2688–2690. [Google Scholar] [CrossRef]
  30. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef]
  31. Yang, Z.H. Maximum-likelihood estimation of phylogeny from DNA sequences when substitution rates differ over sites. Mol. Biol. Evol. 1993, 10, 1396–1401. [Google Scholar]
  32. Mirarab, S.; Bayzid, M.S.; Boussau, B.; Warnow, T. Statistical binning enables an accurate coalescent-based estimation of the avian tree. Science 2014, 346, 1250463. [Google Scholar] [CrossRef] [Green Version]
  33. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. jModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef] [Green Version]
  34. Swofford, D.L. PAUP*: Phylogenetic Analysis Using Parsimony (*and Other Methods), Version 4.0b10; Sinauer Associates: Sunderland, MA, USA, 2002. [Google Scholar]
  35. Ronquist, F.; Teslenko, M. Van der Mark, P.; Ayres, D.L.; Darling, A.; Hohna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
  36. Maddison, M.P.; Maddison, D.R. Mesquite: A Modular System for Evolutionary Analysis. Version 3.5.1. 2018. Available online: http://www.mesquiteproject.org (accessed on 2 November 2020).
  37. Cook, R.E. Growth and development in clonal plant populations. In Population Biology and Evolution of Clonal Organisms; Jackson, J.B.C., Buss, L.W., Cook, R.E., Eds.; Yale University Press: New Haven, CT, USA; London, UK, 1985; pp. 259–296. [Google Scholar]
  38. Rønsted, N.; Law, S.; Thornton, H.; Fay, M.F.; Chase, M.W. Molecular phylogenetic evidence for the monophyly of Fritillaria and Lilium (Liliaceae; Liliales) and the infrageneric classification of Fritillaria. Mol. Phylogenet. Evol. 2005, 35, 509–527. [Google Scholar] [CrossRef]
  39. Chin, H.C.; Chang, M.C.; Ling, P.P.; Ting, C.T.; Dou, F.P. A cytotaxonomic study on Chinese Dioscorea L.—The chromosome numbers and their relation to the origin and evolution of the genus. Acta Phytotaxon. Sin. 1985, 23, 11–18. [Google Scholar]
  40. Wickham, L.D.; Wilson, L.A.; Passam, H.C. The origin, development and germination of bulbils in two Dioscorea species. Ann. Bot. 1982, 50, 621–627. [Google Scholar] [CrossRef]
  41. Ayensu, E.S. Anatomy of the Monocotyledons, VI. Dioscoreales; Clarendon Press: Oxforod, UK, 1972; p. 182. [Google Scholar]
  42. Ronsheim, M.L.; Bever, J.D. Genetic variation and evolutionary trade-offs for sexual and asexual reproductive modes in Allium vineale (Liliaceae). Am. J. Bot. 2000, 87, 1769–1777. [Google Scholar] [CrossRef]
  43. Eckert, C.G.; Lui, K.; Bronson, K.; Corradini, P.; Bruineau, A. Population genetic consequences of extreme variation in sexual and clonal reproduction in an aquatic plant. Mol. Ecol. 2003, 12, 331–344. [Google Scholar] [CrossRef]
  44. Steiner, B.L.; Armbruster, G.F.J.; Scheepens, J.F.; Stöcklin, J. Distribution of bulbil- and seed-producing plants of Poa alpina (Poaceae) and their growth and reproduction in common gardens suggest adaptation to different elevations. Am. J. Bot. 2012, 99, 2035–2044. [Google Scholar] [CrossRef]
  45. Willmer, P. Pollination and Floral Ecology; Priceton University Press: Priceson, NJ, USA, 2011; pp. 1–792. [Google Scholar]
  46. Patiny, S. Evolution of Plant-Pollinator Relationships; Cambrige University Press: Cambridge, UK, 2012; pp. 1–477. [Google Scholar]
  47. Mondo, J.M.; Agre, P.A.; Edemodu, A.; Adebola, P.; Asiedu, R.; Akoroda, M.O.; Asfaw, A. Floral biology and pollination efficiency in yam (Dioscorea spp.). Agriculture 2020, 10, 560. [Google Scholar] [CrossRef]
  48. Weller, S.G.; Sakai, A.K.; Culley, T.M.; Campbell, D.R.; Dunbar-Wallis, A.K. Predicting the pathway to wind pollination: Heritabilities and genetic correlations of inflorescence traits associated with wind pollination in Schiedea salicaria (Caryophyllaceae). J. Evol. Biol. 2006, 19, 331–342. [Google Scholar] [CrossRef]
  49. Iwata, T.; Nagasaki, O.; Ishii, H.S.; Ushimaru, A. Inflorescence architecture affects pollinator behaviour and mating success in Spiranthes sinensis (Orchidaceae). New Phytol. 2012, 193, 196–203. [Google Scholar] [CrossRef]
  50. Weiss, M.R. Floral color changes as cues for pollinators. Nature 1991, 354, 227–229. [Google Scholar] [CrossRef]
  51. Streisfeld, M.A.; Kohn, J.R. Environment and pollinator-mediated selection on parapatric floral races of Mimulus aurantiacus. J. Evol. Biol. 2007, 20, 122–132. [Google Scholar] [CrossRef]
  52. Wang, H.; Conchou, L.; Bessière, J.M.; Cazals, G.; Schatz, B.; Imbert, E. Flower color polymorphism in Iris lutescens (Iridaceae): Biochemical analyses in light of plant-insect interactions. Phytochemistry 2013, 94, 123–134. [Google Scholar] [CrossRef]
  53. Muchhala, N.; Johnsen, S.; Smith, S.D. Competition for hummingbird pollination shapes flower color variation in Andean Solanaceae. Evolution 2014, 68, 2275–2286. [Google Scholar] [CrossRef] [PubMed]
  54. Dellinger, A.S.; Artuso, S.; Pamperl, S.; Michelangeli, F.A.; Penneys, D.S.; Fernández-Fernández, D.M.; Alvear, M.; Almeda, F.; Armbruster, W.S.; Staedler, Y.; et al. Modularity increases rate of floral evolution and adaptive success for functionally specialized pollination systems. Commun. Biol. 2019, 2, 453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Reich, D.; Berger, A.; von Balthazar, M.; Chartier, M.; Sherafati, M.; Schönenberger, J.; Manafzadeh, S.; Staedler, Y.M. Modularity and evolution of flower shape: The role of function, development, and spandrels in Erica. New Phytol. 2020, 226, 267–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Tripp, E.A.; Dexter, K.G.; Stone, H.B. Reproductive character displacement and potential underlying drivers in a species-rich and florally diverse lineage of tropical angiosperms (Ruellia; Acanthaceae). Ecol. Evol. 2021, 11, 4719–4730. [Google Scholar] [CrossRef]
  57. Zhao, Y.M.; Hang, Y.Y.; Zhou, Y.F. The pollination of Dioscorea nipponica ssp. rosthornii. Acta Phytotaxon. Sin. 2008, 17, 15–21. [Google Scholar]
  58. Sadik, S.; Okereke, O.U. Flowering Pollen Grain Germination, Fruiting, Seed Germination and Seedling Development of White Yam, Dioscorea rotundata Poir. Anm. Bot. 1975, 39, 597–604. [Google Scholar] [CrossRef]
  59. Mizuki, I.; Ishida, K.; Kikuzawa, K. Sexual and vegetative reproduction in the aboveground part of a dioecious clonal plant, Dioscorea japonica (Dioscoreaceae). Ecol. Res. 2005, 20, 387–393. [Google Scholar] [CrossRef]
  60. Li, M.M.; Yan, Q.Q.; Sun, X.Q.; Zhao, Y.M.; Zhou, Y.F.; Hang, Y.Y. A preliminary study on pollination biology of three species in Dioscorea (Dioscoreaceae). Life Sci. J. 2014, 11, 436–444. [Google Scholar]
  61. Soto Gomez, M.; Pokorny, M.B.; Kantar, F.; Forest, I.J.; Leitch, B.; Gravendeel, P.; Wilkin, S.; Graham, W.; Viruel, J. A customized nuclear target enrichment approach for developing a phylogenomic baseline for Dioscorea yams (Dioscoreaceae). Appl. Plant Sci. 2019, 7, e11254. [Google Scholar] [CrossRef] [Green Version]
  62. Pei, J.; Ting, C.T.; Chin, H.C.; Su, P.; Tang, S.Y.; Chang, H.C. A preliminary systematic study of Dioscorea L. sect. Stenophora Uline. Acta Phytotaxon. Sin. 1979, 17, 61–72. [Google Scholar]
Figure 1. Characters related to reproduction showing morphological diversity of Dioscorea. (AD) Bulbils. (EH) Floral color. (IK) Perianth opening degree. (LO) Inflorescence architecture. (A) D. kamoonensis. (B) D. melanophyma. (C,N) D. delavayi. (D,K) D. polystachya. (E,O) D. futschauensis. (F) D. zingiberensis. (G) D. yunnanensis. (H) D. subcalva. (I) D. tokoro. (J) D. bulbifera. (L) D.exalata. (M) D. panthaica.
Figure 1. Characters related to reproduction showing morphological diversity of Dioscorea. (AD) Bulbils. (EH) Floral color. (IK) Perianth opening degree. (LO) Inflorescence architecture. (A) D. kamoonensis. (B) D. melanophyma. (C,N) D. delavayi. (D,K) D. polystachya. (E,O) D. futschauensis. (F) D. zingiberensis. (G) D. yunnanensis. (H) D. subcalva. (I) D. tokoro. (J) D. bulbifera. (L) D.exalata. (M) D. panthaica.
Diversity 14 00349 g001
Figure 2. The maximum likelihood tree using the combined data from chloroplast matK, rbcL, trnL-F, psbA-trnH, rpl36-rps8, as well as mitochondria nad1 and rps3. Branches are labeled with maximum likelihood bootstraps higher than 70%, parsimony bootstrap proportions higher than 50% and Bayesian posterior probabilities more than 0.95.
Figure 2. The maximum likelihood tree using the combined data from chloroplast matK, rbcL, trnL-F, psbA-trnH, rpl36-rps8, as well as mitochondria nad1 and rps3. Branches are labeled with maximum likelihood bootstraps higher than 70%, parsimony bootstrap proportions higher than 50% and Bayesian posterior probabilities more than 0.95.
Diversity 14 00349 g002
Figure 3. Ancestral character state reconstruction for bulbils using the parsimony method.
Figure 3. Ancestral character state reconstruction for bulbils using the parsimony method.
Diversity 14 00349 g003
Figure 4. Ancestral character state reconstruction for the inflorescence architecture using the parsimony method.
Figure 4. Ancestral character state reconstruction for the inflorescence architecture using the parsimony method.
Diversity 14 00349 g004
Figure 5. Ancestral character state reconstruction for the floral color using the parsimony method.
Figure 5. Ancestral character state reconstruction for the floral color using the parsimony method.
Diversity 14 00349 g005
Figure 6. Ancestral character state reconstruction for the perianth opening degree using the parsimony method.
Figure 6. Ancestral character state reconstruction for the perianth opening degree using the parsimony method.
Diversity 14 00349 g006
Table 1. Taxa used in this study with locality, voucher information and GenBank accession numbers.
Table 1. Taxa used in this study with locality, voucher information and GenBank accession numbers.
SpeciesLocalityVoucherGenBank Accession No.
matKtrnL-FrbcLpsbA-trnHrpl36-rps8nad1rps3
Dioscorea nipponica MakinoLin’an, Zhejiang, China NAS 0648570 AY957600 *DQ841308AF307455 *GQ265171GQ265222GQ265123GQ265268
D. nipponica subsp. rosthornii (Prain & Burkill) C. T. Ting Tianshui, Gansu, China NAS 0648571DQ974184DQ841309DQ408178GQ265172GQ265223GQ265122GQ265269
D. althaeoides R. KnuthDêqên, Yunnan, China NAS 0648572 EU407548EU301741EU407550GQ265182GQ265233GQ265135GQ265281
D. tokoro MakinoAnhua, Hunan, China NAS 0648573 DQ974186DQ841312DQ408180GQ265174GQ265225GQ265125GQ265271
D. zingiberensis C. H. Wright Mt. Hengshan, Hunan, China NAS 0646476 AY973831 *DQ841318AY939889 *GQ265154GQ265206GQ265105GQ265251
D. sinoparviflora C. T. Ting, M. G. Gilbert & N. J. Turland Lijiang, Yunnan, China NAS 0648574DQ974179DQ841326DQ408171GQ265163GQ265212GQ265112GQ265258
D. deltoidea Wall. ex Griseb. Kunming, Yunnan, China NAS 0648575EF614207DQ841305EF614218GQ265169GQ265220GQ265120GQ265266
D. panthaica Prain & BurkillLijiang, Yunnan, ChinaY. F. Zhou & B. C. Wu 200308015GQ265088GQ265291GQ265187GQ265184GQ265235GQ265136GQ265283
D. biformifolia C. Pei & C. T. TingMt. Eshan, Yunnan, China NAS 0648576EU407549EU301742EU301740GQ265147GQ265196GQ265097GQ265243
D. gracillima Miq. Mt. Lushan, Jiangxi, ChinaNAS 0648577DQ974190DQ841315DQ408164GQ265150GQ265201GQ265101GQ265247
D. collettii Hook. f. var. collettiiJinghong, Yunnan, China NAS 0648578DQ974178DQ841300DQ408173GQ265141GQ265192GQ265093GQ265239
D. collettii var. hypoglauca (Palibin) C. Pei & C. T. TingMt. Hengshan, Hunan, China NAS 0648579DQ974176DQ841319EF614220GQ265155GQ265205GQ265106GQ265252
D. futschauensis Uline ex R. KnuthYongtai, Fujian, ChinaNAS 0648580DQ974175DQ841316DQ408166GQ265151GQ265202GQ265102GQ265248
D. spongiosa J. Q. Xi, M. Mizuno & W. L. ZhaoMt. Hengshan, Hunan, ChinaNAS 0648581 DQ974191DQ841317DQ974194GQ265153GQ265204GQ265104GQ265250
D. banzhuana C. Pei & C. T. Ting Mengzi, Yunnan, China NAS 0648582 DQ974182DQ841301DQ408174GQ265167GQ265218GQ265118GQ265264
D. simulans Prain & BurkillGuilin, Guangxi, China NAS 0648583 EF614206DQ841320EF614217GQ265138GQ265189GQ265090GQ265236
D. esculenta (Lour.) BurkillLingshui, Hainan, ChinaNAS 0648585AY956497 *DQ841298AY904794 *GQ265180GQ265231GQ265131GQ265277
D. esculenta var. spinosa (Roxburgh ex Prain & Burkill) R. KnuthLingshui, Hainan, ChinaB. C. Wu200804023GQ265087GQ265290GQ265186GQ265181GQ265232GQ265134GQ265280
D. tentaculigera Prain & Burkill Lincang, Yunnan, China NAS 0648465GQ265089GQ265292GQ265188GQ265183GQ265234GQ265137GQ265282
D. yunnanensis Prain & Burkill Lijiang, Yunnan, China NAS 0648459EF614209GQ265288EF614221GQ265161GQ265213GQ265113GQ265259
D. subcalva Prain & Burkill Tianlin, Guangxi, China NAS 0648544EF614208EF614222EF614214GQ265160GQ265211GQ265111GQ265257
D. subcalva var. submollis (R. Knuth) C. T. Ting & P. P. Ling Mt. Jinfo, Chongqing, ChinaNAS 0648545EF614204GQ265287EF614216GQ265152GQ265203GQ265103GQ265249
D. nitens Prain & Burkill Lijiang, Yunnan, China NAS 0648586EF614205EF614223EF614215GQ265162GQ265214GQ265114GQ265260
D. bulbifera L.Jinghong, Yunnan, China NAS 0648587 AY956488 *EF619352AY904791 *GQ265178GQ265229GQ265132GQ265278
D. melanophyma Prain & BurkillMengzi, Yunnan, China NAS 0648548 EF614210DQ841303DQ408176GQ265143GQ265194GQ265095GQ265241
D. kamoonensis KunthMengzi, Yunnan, ChinaNAS 0648549 EF028332DQ841302DQ408175GQ265142GQ265193GQ265094GQ265240
D. delavayi Franchet Kunming, Yunnan, China NAS 0648550GQ265085GQ265284DQ974196GQ265148GQ265199GQ265100GQ265246
D. menglaensis H. Li Jinghong, Yunnan, China NAS 0648461GQ265086GQ265285GQ265185GQ265146GQ265198GQ265099GQ265245
D. pentaphylla LinnaeusGuilin, Guangxi, China NAS 0648551AY972483 *DQ841327AF307470 *GQ265140GQ265191GQ265092GQ265237
D. esquirolii Prain & BurkillLongzhou, Guangxi, China NAS 0648552DQ974177DQ841322DQ408168GQ265139GQ265190GQ265091GQ265238
D. hispida Dennstedt.Longzhou, Guangxi, China NAS 0648553AY957589 *DQ841323AF307463 *GQ265145GQ265197GQ265098GQ265244
D. aspersa Prain & Burkill Mengzi, Yunnan, China NAS 0648588 EF614211DQ841304EF614213GQ265168GQ265219GQ265119GQ265265
D. polystachya TurczaninowJurong, Jiangsu, ChinaNAS 0648589 EF028331DQ841313DQ408181GQ265144GQ265195GQ265096GQ265242
D. japonica ThunbergLin’an, Zhejiang, ChinaNAS 0648590DQ974183DQ841307AF307457 *GQ265170GQ265221GQ265121GQ265267
D. cirrhosa Loureiro Longzhou, Guangxi, ChinaNAS 0648591 EF028329DQ841324AY904792 *GQ265158GQ265209GQ265109GQ265255
D. cirrhosa var. cylindrica C. T. Ting & M. C. Chang Mt. Diaoluo, Hainan, ChinaNAS 0648592 DQ974189DQ841314DQ408184GQ265179GQ265230GQ265133GQ265279
D. wallichii J. D. Hooker--AY973830 *-AY939888 *----
D. glabra RoxburghLongzhou, Guangxi, ChinaNAS 0648593 AY956501 *DQ841321AF307456*GQ265157GQ265208GQ265108GQ265254
D. fordii Prain & Burkill Guilin, Guangxi, China NAS 0648594 EF028333DQ841299DQ974195GQ265156GQ265207GQ265107GQ265253
D. persimilis Prain & BurkillMingxi, Fujian, China NAS 0648595DQ974193DQ841328DQ408165GQ265175GQ265226GQ265127GQ265273
D. exalata C. T. Ting & M. C. Chang Tianlin, Guangxi, ChinaNAS 0648596EF028330DQ841325DQ408170GQ265159GQ265210GQ265110GQ265256
D. alata LinnaeusJinghong, Yunnan, ChinaNAS 0648597 AB040208 *DQ841331AY667098 *GQ265165GQ265216GQ265116GQ265262
D. decipiens J. D. Hooker Jinghong, Yunnan, ChinaNAS 0648598DQ974181DQ841329AF307454 *GQ265166GQ265217GQ265117GQ265263
D. composita Hemsl.Jinghong, Yunnan, ChinaNAS 0648405 DQ974180DQ841330DQ408172GQ265164GQ265215GQ265115GQ265261
D. sansibarensis PaxBotanical Garden Regen GermanyY. F. Zhou200403004DQ974187DQ841296AY939883 *GQ265177GQ265228GQ265129GQ265275
D. caucasica LipskyLyon, France NAS 0648584 DQ974188DQ841297DQ408182--GQ265130GQ265276
D. elephantipes Engl.South Africa N. Sheng200511014AY956496 *DQ841306AF307461 *GQ265176GQ265227GQ265128GQ265274
D. villosa L.AmericaNAS 0648463 -GQ265286DQ006092 *GQ265149GQ265200GQ265126GQ265272
Tacca chantieri André--AY973837 *FJ194472 *AJ235810 *EF590744 *-DQ786152 *-
* Sequences obtained from Genbank.
Table 2. Markers and primers used in this study.
Table 2. Markers and primers used in this study.
MarkerPrimer NameDirectionPrimer Sequence (5′ to 3′)
matKmatK MFforwardATT TGC GAT CTA TTC ATT CAA T
matK MRreverseTGA GAT TCC GCA GGT CAT T
rbcLrbcL m3forwardTAT CTT AGC GCC ATT CCG AGT A
rbcL m4reverseCGC GGA TAA TTT CAT TAC CTT C
trnL-FtrnL-F cforwardCGA AAT CGG TAG ACG CTA CG
trnL-F freverseATT TGA ACT GGT GAC ACG AG
psbA- trnHpsbA F1forwardAAT GCT CAC AAC TTY CCT CTA
trnH R1reverseCCA CTG CCT TGA TCC ACT TG
rpl36- rps8rpl36 F1forwardTTA CCC YTG TCT YTG TTT ATG
rps8 R1reverseCTA CGA GAR GGT TTT ATT GAA
nad1nad1 F1forwardCCT TGT GAG CAC GTT TGG AT
nad1 R1reverseGAC AAT CTC ACT CGA ATT ACA G
rps3rps3 F1forwardGTT CGA TAC GTC CAC CTA C
rps3 R1reverseGTA CGT TTC GGA TAT RGC AC
Table 3. Data matrix of morphological characters used in this study.
Table 3. Data matrix of morphological characters used in this study.
SpeciesCharacter
abcd
Dioscorea sect. StenophoraD. nipponica0110
D. nipponica subsp. rosthornii0210
D. althaeoides0310
D. tokoro0310
D. zingiberensis0100
D. sinoparviflora0100
D. deltoidea0110
D. panthaica0310
D. biformifolia0310
D. gracillima0110
D. collettii0110
D. collettii var. hypoglauca0110
D. futschauensis0300
D. spongiosa0310
D. banzhuana0310
D. simulans0200
D. caucasica0110
D. villosa0111
D. sect. CombiliumD. esculenta0110
D. esculenta var. spinosa0110
D. sect. ShannicoreaD. tentaculigera1110
D. yunnanensis0110
D. subcalva0110
D. subcalva var. submollis0110
D. nitens0110
D. sect. OpsophytonD. bulbifera1101
D. sect. BotryosicyosD. melanophyma1212
D. kamoonensis12 12
D. delavayi1212
D. menglaensis1312
D. pentaphylla1311
D. esquirolii1211
D. sect. LasiophytonD. hispida0312
D. sect. EnantiophyllumD. aspersa011 2
D. polystachya111 2
D. japonica1112
D. cirrhosa111 2
D. cirrhosa var. cylindrica131 2
D. wallichii031 2
D. glabra131 2
D. fordii131 2
D. persimilis131 2
D. exalata131 2
D. alata131 2
D. decipiens1312
D. sect. ApodostemonD. composita0111
D. sect. MacrouraD. sansibarensis1112
D. sect. TestudinariaD. elephantipes0200
outgroupTacca chantieri0000
Table 4. Informative parameters for the seven molecular markers.
Table 4. Informative parameters for the seven molecular markers.
DNA RegionAligned Length (bp)Variable Site
(bp/%)
Informative Site
(bp/%)
Tree LengthCIRI
matK1032237/23.0163/15.83170.8520.959
rbcL1142134/11.784/7.42200.6640.867
trnL-F939280/29.880/8.53850.8160.905
psbA-trnH344113/32.882/23.81600.8500.955
rpl36-rps8857111/13.064/7.51420.8170.949
nad11478207/14.0125/8.52460.9020.949
rps3142675/5.319/1.3890.8760.929
7 DNA72181157/16.0617/8.516400.7820.913
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, M.; Sun, X.; Xue, J.-Y.; Zhou, Y.; Hang, Y. Evolution of Reproductive Traits and Implications for Adaptation and Diversification in the Yam Genus Dioscorea L. Diversity 2022, 14, 349. https://doi.org/10.3390/d14050349

AMA Style

Chen M, Sun X, Xue J-Y, Zhou Y, Hang Y. Evolution of Reproductive Traits and Implications for Adaptation and Diversification in the Yam Genus Dioscorea L. Diversity. 2022; 14(5):349. https://doi.org/10.3390/d14050349

Chicago/Turabian Style

Chen, Min, Xiaoqin Sun, Jia-Yu Xue, Yifeng Zhou, and Yueyu Hang. 2022. "Evolution of Reproductive Traits and Implications for Adaptation and Diversification in the Yam Genus Dioscorea L." Diversity 14, no. 5: 349. https://doi.org/10.3390/d14050349

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop