Skip to main content
Advertisement
Browse Subject Areas
?

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

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Diversification of Angraecum (Orchidaceae, Vandeae) in Madagascar: Revised Phylogeny Reveals Species Accumulation through Time Rather than Rapid Radiation

  • Herinandrianina N. Andriananjamanantsoa ,

    Contributed equally to this work with: Herinandrianina N. Andriananjamanantsoa, Shannon Engberg, Edward E. Louis Jr, Luc Brouillet

    ratahx@yahoo.fr

    Affiliation Département de sciences biologiques, Université de Montréal, Montréal, Québec, Canada

  • Shannon Engberg ,

    Contributed equally to this work with: Herinandrianina N. Andriananjamanantsoa, Shannon Engberg, Edward E. Louis Jr, Luc Brouillet

    Affiliation Omaha’s Henry Doorly Zoo and Aquarium, Omaha, Nebraska, United States of America

  • Edward E. Louis Jr ,

    Contributed equally to this work with: Herinandrianina N. Andriananjamanantsoa, Shannon Engberg, Edward E. Louis Jr, Luc Brouillet

    Affiliation Omaha’s Henry Doorly Zoo and Aquarium, Omaha, Nebraska, United States of America

  • Luc Brouillet

    Contributed equally to this work with: Herinandrianina N. Andriananjamanantsoa, Shannon Engberg, Edward E. Louis Jr, Luc Brouillet

    Affiliation Département de sciences biologiques, Université de Montréal, Montréal, Québec, Canada

Abstract

Angraecum is the largest genus of subtribe Angraecinae (Orchidaceae) with about 221 species. Madagascar is the center of the diversity for the genus with ca. 142 species, of which 90% are endemic. The great morphological diversity associated with species diversification in the genus on the island of Madagascar offers valuable insights for macroevolutionary studies. Phylogenies of the Angraecinae have been published but a lack of taxon and character sampling and their limited taxonomic resolution limit their uses for macroevolutionary studies. We present a new phylogeny of Angraecum based on chloroplast sequence data (matk, rps16, trnL), nuclear ribosomal (ITS2) and 39 morphological characters from 194 Angraecinae species of which 69 were newly sampled. Using this phylogeny, we evaluated the monophyly of the sections of Angraecum as defined by Garay and investigated the patterns of species diversification within the genus. We used maximum parsimony and bayesian analyses to generate phylogenetic trees and dated divergence times of the phylogeny. We analyzed diversification patterns within Angraecinae and Angraecum with an emphasis on four floral characters (flower color, flower size, labellum position, spur length) using macroevolutionary models to evaluate which characters or character states are associated with speciation rates, and inferred ancestral states of these characters. The phylogenetic analysis showed the polyphyly of Angraecum sensu lato and of all Angraecum sections except sect. Hadrangis, and that morphology can be consistent with the phylogeny. It appeared that the characters (flower color, flower size, spur length) formerly used by many authors to delineate Angraecum groups were insufficient to do so. However, the newly described character, position of the labellum (uppermost and lowermost), was the main character delimiting clades within a monophyletic Angraecum sensu stricto. This character also appeared to be associated with speciation rates in Angraecum. The macroevolutionary model-based phylogeny failed to detect shifts in diversification that could be associated directly with morphological diversification. Diversification in Angraecum resulted from gradual species accumulation through time rather than from rapid radiation, a diversification pattern often encountered in tropical rain forests.

Introduction

Madagascar is known for its rich biodiversity [1] and is a focus for evolutionary biologists who study the causes of species diversification. Many studies have been conducted using the high endemicity and diversity found in this isolated island as a model of diversification processes in various taxa [210]. Most of these studies came to the conclusion that species diversification resulted from rapid radiation. High morphological variation associated with low genetic divergence appeared to be a signature of the diversification processes operating on the island of Madagascar [8] as well as on many other islands such as New Zealand, Hawaii, or those of the Caribbean [1113]. With the improvement of methods for macroevolutionary studies [1418] it has been advocated that species diversification is not necessarily linked with adaptive radiations [19], but could be a result of a gradual accumulation of ancestral lineages through time [20]. The problems associated with studies on diversification are primarily due to insufficient data, such as incomplete taxon sampling or phylogenetic uncertainties [17,21]. Macroevolutionary studies require a good knowledge of paleontological events and eventually a fossil record to help calibrate the phylogenies, but most of the time these are lacking, making the interpretation of evolutionary histories controversial. Understanding the diversification processes of a group of organisms could be useful for biodiversity conservation [22], especially for hotspots like Madagascar [23] where priority should be given to the most valuable species due to a lack of conservation funds.

The genus Angraecum is the second largest group of orchids in Madagascar; it has a high morphological diversity, which makes it a good candidate for macroevolutionary studies [21]. Angraecum includes approximately 221 species [24,25] distributed from sub-Saharan Africa to Madagascar, the Indian Ocean Islands (IOI: Comoros, Mauritius, Reunion, Rodrigues, and Seychelles), and Sri Lanka. Madagascar is the center of diversity for the genus with ca. 142 species, of which 90% are endemic. The majority of species are epiphytic, but some are lithophytic. Epiphytic plants are found in the tropical rain forest across the eastern slopes of mountains or in mesic forests in the central highland, while lithophytic species are found on inselbergs or on limestone. Recent molecular phylogenetic work [24] placed Angraecum in subtribe Angraecinae, the latter sister to subtribe Aeridinae of tribe Vandeae. A phylogenetic reconstruction of the tribe Vandeae [26] revealed the polyphyly of Angraecinae sensu Dressler [27] and of many of its genera including Angraecum [26,28]. Accordingly, a new circumscription has been proposed by merging Angraecinae and Aerangidinae into a single subtribe Angraecinae [28], which includes 47 genera and ca. 762 species [24]. Even though revisions were made to resolve the polyphyly of several genera (e.g. [29,30]), similar attempts with Angraecum proved difficult because of limited sampling [26,28]. In general, research on Angraecum is complicated by the large number of species and great morphological diversity within the genus, with morphological similarities to taxa in other genera of subtribe Angraecine confusing matters further. Garay [31] proposed 19 sections to accommodate the species according to morphological descriptions (essentially floral, e.g. flower color, flower size, spur length), as was done by previous authors (e.g. [3235]). Recent molecular phylogenic work revealed that Garay’s sections are polyphyletic [28,36]. Micheneau et al. [28] showed that the morphology-based classification did not reflect the molecular phylogeny, and concluded that most of the sections of Garay [31] were non-monophyletic. This study was based on Mascarene species and was lacking samples from Madagascar. Some authors proposed to remove all species that cause the polyphyly with the aim of making Angraecum monophyletic [3638]. Thus, all strictly African sections were removed from Angraecum sensu lato [38] and elevated to generic rank (Angraecoides, Dolabrifolia, Eichlerangraecum and Pectinariella), making the taxa sampled monophyletic, but as mentioned above, the study was not comprehensive, lacking Malagasy taxa.

Although the evolution of pollinia, the adaptation to epiphytic habitat and the development of crassulacean acid metabolism (CAM) photosynthesis was shown to be involved in the earlier diversification of the Orchidaceae [39], little is known about the relationship between morphological variation and species diversification. With the morphological diversity observed in Angraecum, we were interested in testing whether the characters historically used to define the sections, essentially flower color, flower size and spur length, were phylogenetically useful and involved in species diversification. Questions also arose on how the diversification of the genus occurred. Here, we present a phylogenetic reconstruction of Angraecum sensu Garay based on molecular DNA sequence data and a more comprehensive sample of species. Using this phylogenetic framework combined with morphological data, we attempt to resolve the systematic problems existing at the sectional level, and try to understand the patterns of species diversification in Angraecum. Therefore, the main objectives of the current study are to (1) reconstruct the phylogeny of Angraecum sensu Garay [31] using a larger sample of Malagasy species and most available Angraecinae species; (2) test the monophyly of morphologically defined sections; (3) evaluate whether species diversification follows morphological diversification; (4) assess the validity of the morphological characters for taxonomy; and (5) to establish whether diversification was the result of rapid radiation or more gradual taxon accumulation. To avoid confusion, we will use Angraecum to designate the sensu stricto genus, and Angraecum sensu lato to identify the widest concept of the genus as defined by Garay.

Materials and Methods

Taxon sampling and study site

Plant tissue for DNA extraction was obtained from field collected silica gel dried samples [40] of 36% of Malagasy Angraecum s.l. and Angraecinae species. The remaining materials, in the form of sequences data, came from previous studies [26,28,36,41]. A total of 194 specimens, of which 69 were newly sampled, were included in our analyses, comprising 98 Angraecum s.l., 17 Jumellea, 10 Aeranthes, and 40 other genera (S1 Table). Acampe ochracea, Aerides odorata, Phalaenopsis cornu-cervi, Vanda tricolor (subtribe Aeridinae), and Polystachya fulvilabia (subtribe Polystachyinae) were used as outgroups based on previous phylogenetic studies [24,26,28].

The study focused on Madagascar as this is the center of diversity of Angraecum [42]. Sampling was carried out from 2007 to 2012 at 11 sites. The characteristics of each site are shown in Table 1 where evergreen and deciduous formations correspond to the description of Du Puy & Moat [43]. The expeditions were carried out in collaboration with Madagascar Biodiversity Partnership, the University of Antananarivo, the Ministry of Environment, Ecology and Forest, and Madagascar National Park. Research permits were issued by the Ministry of Environment, Ecology and Forest in agreement with Madagascar National Park and the University of Antananarivo. Field studies did not involve endangered or protected species.

Because of conservation policy in Madagascar National Parks, the collection of herbarium specimens was not allowed in protected areas. Therefore, pictures were taken to serve as vouchers and were stored at the Marie-Victorin Herbarium (MT), which will be accessible in the future via Canadensys (http://data.canadensys.net/explorer/en/search). Elsewhere, voucher specimens were collected and deposited at the national herbarium of Madagascar (TAN) and MT. Overall, 32 of the 69 sampled specimens were vouchered (S1 Table).

Morphological data

To investigate morphological evolution and diversification patterns, 39 characters (13 vegetative and 26 floral) were scored for the taxa represented in our molecular sampling. Type specimens preserved in the herbaria at Kew (K) and Paris (P), where over 80% of holotypes are located, were scored for each Angraecum s.l. species. Missing characters were documented from literature descriptions. For type specimens preserved in other herbaria (e.g. B, BM, MO) that we did not borrow and for other genera, characters were scored from photographs of living material and voucher specimens, as well as from literature descriptions [34,42,44]. Character descriptions and other background information are provided in S2 Table, while the generated matrix is presented in S3 Table. The independence of the morphometric characters was not assessed, as we did not anticipate this would unduly impact on the results.

PCR amplification and DNA sequencing

Total DNA was extracted from 20–30 mg of silica-dried gel leaf material following the modified hexadecylmethylammonium bromide (2x CTAB, 2% (w/v)) extraction protocol of Doyle and Doyle [45]; 1% polyvinylpyrrolidone (PVP) and 0.2% of β-mercaptoethanol was added to the total volume of the extraction buffer. Three plastid DNA markers were amplified: matK coding gene, rps16 intergeneric spacer, and trnL intron. The amplification of the matK region was performed using the barcoding primers 472F/1248R designed by Yu et al. [46]. The rps16 region was amplified using the primers 1F/2R designed by Oxelman et al. [47]. The trnL intron was amplified using the primers 49873F/50272R designed by Taberlet et al. [48]. The nuclear ribosomal internal transcribed spacer ITS2 region was amplified using the primers S2F/S3R designed by Chen et al. [49]. The PCR reactions contained, 1X PCR reaction buffer, 2.5 mM MgCl2, 0.16 μM of each primer, 0.2 mM of each dNTP, 0.4% bovine serum albumin (BSA), 2 units of Taq DNA polymerase, 30 ng of template DNA, adjusted to a final volume of 25 μL with de-ionized water. PCR conditions are the same as described in Yu et al. [46]. PCR amplifications were performed on a GeneAmp PCR System 9700 thermocycler (Applied Biosystems, Foster City, CA), and resulting PCR products were purified using exonuclease I and shrimp alkaline phosphatase (ExoSAP; [50]). The purified products were cycle sequenced using a BigDye® terminator sequencing kit (Life Technologies, Carlsbad, CA). Sequences were analyzed with an Applied Biosystems 3130xl genetic analyzer at the Omaha’s Henry Doorly Zoo and Aquarium (NE, USA). Sequence fragments were aligned to generate a consensus sequence using Sequencher® 4.10 (Gene Codes Corporation; Ann Arbor, MI). Sequence data were also obtained from the Canadian Center for DNA Barcoding (part of matK and ITS2) and from GenBank for previous studies [26,28,36,41]. Following automatic alignment using SeaView [51], alignments were edited manually using BioEdit [52]. All newly generated sequences have been deposited in BOLD and GenBank (S1 Table), and the matrices in TreeBase (http://purl.org/phylo/treebase/phylows/study/TB2:S19429).

Phylogenetic analyses

Five matrices were produced: [1] combined plastid, [2] nuclear ribosomal, [3] morphological, [4] combined plastid and morphological, and [5] combined molecular (plastid and nuclear ribosomal) and morphological. They were analyzed using maximum parsimony (MP) and Bayesian analyses (BA). Tree searches under parsimony were conducted in PAUP* [53]. A preliminary heuristic search was performed with 1000 replicates of random addition sequence, tree bisection–reconnection (TBR) branch swapping, retaining twenty most parsimonious trees at each replicate. Starting with the trees kept in memory from this initial analysis, a second heuristic search was performed with TBR and 10000 trees were saved. A strict consensus tree was constructed for each analysis. Branch support was estimated using 5000 bootstraps replicates under a heuristic strategy with one random addition-sequence replicate TBR branch swapping. Bayesian analyses were performed with MrBayes [54]. The best nucleotide substitution model was selected with jModelTest2 [55] using the Akaike information criteria [56]. For all regions, the GTR+I+G model scored best and was selected. For the combined molecular and morphological matrix, data were partitioned as DNA and standard respectively. Two parallel runs of eight Metropolis Coupled Markov Chain Monte Carlo (MCMCMC) each, and four swaps per swapping cycles for 15 million generations were undertaken. Trees were sampled every 1000 generations, and the first 25% generations were discarded as burnin. The 50% majority consensus tree with Bayesian clade credibility was built from post-burnin trees.

Estimation of divergence times

Divergence times were estimated using a relaxed molecular clock approach as implemented in BEAST [57]. The combined plastid matrix was used as input data in BEAUti. The GTR+G+I model was selected as substitution model. A relaxed lognormal molecular clock model was selected. The Yule model was selected as tree prior. The age for the root of the tree was set to a normal distribution with mean 35 million years (Ma) and a standard deviation of 3 (giving a 95% CI ranging from 30.07–39.93 Ma). Because fossil data are rare in the Orchidaceae, only three fossils having been recorded so far [58], none of them close to our group, we used the age estimate of Phalaenopsis [59], a member of subtribe Aeridinae sister to Angraecinae which occupies a phylogenetic position near the base of tribe Vandeae [24,60], to calibrate the stem root. The prior distribution of the ‘ucld.mean’ parameter was set to an exponential distribution (mean = 10.0, initial value = 1.0). Four separate runs were performed in BEAST with 50 million generations each, sampling parameters and trees every 1000 generations. Trees were summarized with burnin values set to the first 25% of trees sampled using TreeAnnotator and were summarized in a maximum clade credibility tree.

Diversification analyses

In order to assess diversification patterns, we evaluated the state-dependent diversification of morphological characters using the Binary State Speciation and Extinction (BiSSE) and the MultiState Speciation and Extinction (MuSSE) models implemented in the R package ‘diversitree’ [61], and the speciation/extinction and phenotypic/evolution models using BAMM [62]. Four floral characters were chosen for these analyses because of their taxonomic interest: flower colors, flower size, spur length, and labellum position (S2 Table). Three of these characters (flower color, flower size and spur length) have been used previously to delineate sections in Angraecum (e.g. [31,34]), while the labellum position was added because it appeared useful for sectional delimitation. Since these methods require ultrametric and fully bifurcating trees, we used the BEAST maximum clade credibility tree as input.

Two characters were analyzed using BiSSE: flower color and labellum position. Since we were interested in the effect the green and white flower colors may have on diversification, the taxa that did not fit in these two colors (for instance, pink or purple flowered taxa) were excluded from the analyses using the ‘drop.tip2’ function of the R package ‘phyloch’ [63]. Four models were tested: (M1) a full model that allows all parameters' (λ: speciation rate, μ: extinction rate, q: transition rate) to vary, (M2) a constrained model allowing speciation and transition rates to vary while keeping extinction rates equal between states (μ0 ~ μ1), (M3) a constrained model allowing extinction and transition rates to vary while keeping speciation rates equal between states (λ0 ~ λ1), and (M4) a constrained model allowing speciation and extinction rates to vary while keeping transition rates equal between states (q0 ~ q1). Two characters were analyzed with MuSSE: flower size and spur length. To categorise these continuous characters, partition analyses using K-means were conducted (S1 Fig and S4 Table) as implemented in the R package ‘stats’ [64,65]. Both characters were treated as ordered. A preliminary run was performed to determine the best transition model. Only transitions that we considered biologically plausible were tested (e.g.: transitions between very small flower to very large flower, or between very short spur to very long spur are less probable). For flower size, which has four character states (1: small, 2: medium, 3: large, 4: very large), three models were tested: (M1) a full model that allows all parameters (λ, μ, q) to vary, (M2) a constrained model where character evolution was possible in both directions between neighboring states (1 ↔ 2 ↔ 3 ↔ 4), and (M3) a constrained model where evolution could be either unidirectional or bidirectional between neighboring states (1 ←2 ←3 ↔ 4). For spur length, which has four character states (1: short, 2: medium, 3: long, 4: very long), three models were also tested: (M1) a full model that allows all parameters (λ, μ, q) to vary, (M2) a constrained model where character evolution was bidirectional and possible only among neighboring states (1 ↔ 2 ↔ 3 ↔ 4), and (M3) a constrained model where character evolution could be either unidirectional or bidirectional between states that are not necessarily adjacent (reticulated) (1 ←2 ↔ 3 ↔ 4, 1←3, 1←4). Having established the best fit transition model, we used it as a constraint while testing the three models of diversification: (M1) an unconstrained model that allowed all parameters (λ, μ, q) to vary, (M2) a constrained model allowing speciation and transition rates to vary while keeping extinction rates equal between states, and (M3) a constrained model allowing extinction and transition rates to vary while keeping speciation rates equal between states. After testing the models with BiSSE and MuSSE, the posterior probabilities of the parameters were computed under a Bayesian framework after setting the priors to be exponential. The MCMC was run for 10000 generations sampling parameters every 100 generations, and the posterior distributions of parameters were summarized using the function ‘profiles.plot’ implemented in ‘diversitree’ [61].

Many criticisms have been made concerning the SSE family (BiSSE, MuSSE, GeoSSE, etc.) type I error rate and limitations [66,67]. It has been reported that “within-clade pseudoreplications” might result in erroneously significant results [66]. The use of additional methods like BAMM has been proposed to reinforce the results implemented under SSE’s [67]. Both methods use different approaches but are complementary: MuSSE looks at the effects of character states on diversification, while BAMM quantifies diversification rates and detects macroevolutionary rate shifts events (speciation, extinction and phenotypic evolution) across phylogenetic trees.

For the speciation/extinction model implemented in BAMM, we estimated our total sampling at 40% and we fractioned the data according to the number of species per genus being represented in our phylogeny (S4 Table) in order to reduce the bias in estimating parameters under an assumption of incomplete sampling [62]. To reduce the weight of the outgroup taxa we excluded them from the analyses using the ‘drop.tip2’ function of the R package ‘phyloch’ [63]. For the phenotypic/evolution model, we assessed the regime of morphological evolution of two of the four characters mentioned above, flower size and spur length (S4 Table), since this model only treats continuous characters. For each model (speciation/extinction and phenotypic/evolution), BAMM was run for 5,000,000 generations and parameters were sampled every 1000 generations. The parameter priors were set using parameters generated from the R package ‘BAMMtools’ [18]. The rate shift configurations and the rate through-time generated from ‘bammdata outputs’ were analyzed using the R package ‘BAMMtools’.

Character state reconstruction

We examined the character evolution of four floral characters previously used in taxonomy (flower colors, flower size, spur length, and labellum position) using the Markov discrete character evolution [14] as implemented in the R package ‘diversitree’ [61]. Because this method requires ultrametric and fully bifurcating trees, we used the BEAST tree as input. No constraint was applied to the analyses leaving all parameters free.

Results

Phylogenetic relationships

Of the 69 newly sampled specimens that we sequenced, all were fully amplified with matK for a total length of 942 base pairs (bp); 66 with rps16 (1192 bp) except Angraecum sterophyllum, A. rhynchoglossum and Lemurella papillosa which failed to amplify; and 65 with trnL (1553 bp) except A. pseudofilicornu, A. rhynchoglossum, Oeoniella polystachys and Oeonia rosea. Only 13 of 69 samples were sequenced with ITS2 for a total length of 381 bp. When combined with GenBank sequences data, each individual matrix was composed of 190 taxa for matK, 144 for rps16, 170 for trnL and 88 for ITS2. Because of the small amount of samples available with ITS2, we decided to include only taxa present in this matrix to run the combined molecular (plastid and nuclear) and morphological analyses for a total of 88 taxa and 4107 characters. Therefore, the combined plastid matrix contained 3687 characters and 194 taxa. Sequences that were unavailable were treated as missing data. For the trnL intron, the ambiguous variable sites (360 bp) were excluded from our analyses. For the whole data set, 2361 characters were constant, 467 (13.8%) variable characters were parsimony-uninformative, and 537 (16%) characters were parsimony-informative. The combined plastid matrix produced five trees of 2202 steps with a consistency index (CI) of 0.56 and a retention index (RI) of 0.77.

The MrBayes, BEAST, and MP analyses were congruent. The 50% majority-rule consensus tree from Bayesian analyses of the combined plastid matrix is displayed in Fig 1. Our results support a monophyletic subtribe Angraecinae (PP 1.0, BP 100). Two well-supported clades were identified within Angraecinae: clade I (PP 1.0, BP 77) comprised of Malagasy, IOI, African and American genera, and clade II (PP 1.0, BP 91) with Malagasy and IOI genera. Clade II had more branch support and showed more resolution than clade I. Two main subclades are observed within clade I: a Malagasy-IOI clade A (PP 1.0), and an African-American clade B (PP 1.0). Four Malagasy Angraecum s.l. species (A. perparvulum, A. cf. humile, A. pterophyllum, and A. rhynchoglossum) are nested within clade A. Three major subclades were observed in clade II: Aeranthes (PP 1.0, BP 98), Jumellea (PP 1.0, BP 97), and Angraecum (PP 1.0). Here, we define Angraecum as a monophyletic group including all Angraecum species and all other taxa nested within it in clade II. Based on branch support (PP, BP) and morphological resemblance, eleven clades are observed within Angraecum (Fig 1, clade C to M). From the base to the top of the tree are: clade C (PP 1.0, BP 74) comprised members of sections Pectinaria, Pseudojumellea, Arachnangraecum and Filangis; clade D (PP 1.0, BP 81) comprising member of sections Perrierangraecum, Angraecum, Arachnangraecum and Filangis; clade E (PP 1.0) which includes A. sesquipedale, and A. sororium of section Angraecum; clade F (PP 1.0, BP 100) section Hadrangis; clade G (PP 1.0, BP 61) with section Humblotiangraecum and a member of section Perrierangraecum; clade H (PP 1.0, BP 94) section Boryangraecum; clade I (PP 1.0, BP 100) section Angraecoides; clade J section Arachnangraecum; clade K (PP 1.0, BP 98) section Angraecum (A. eburneum); clade L (PP 1.0, BP 75) comprised of members of sections Angraecum and Pseudojumellea; clade M (PP 1.0, BP 57) composed of sections Acaulia, Boryangraecum, Chlorangraecum, Gomphocentrum, Lemurangis, and Lepervenchea. Clades E and J received weak support despite the fact that their positions were supported in a strict consensus tree. Two species are not included in any of the clades, A. nanum sister to clade I (PP 1.0, BP 72), and A. amplexicaule intermediate between clade L and M (Fig 1). Furthermore, two Malagasy genera, Oeoniella and Sobennikoffia, are nested within Angraecum. Even though many clades were strongly supported within Angraecum, resolution between clades was low.

thumbnail
Fig 1. Phylogenetic relationships within subtribe Angraecinae.

50% Bayesian majority-rule consensus tree from combined plastid data (matK, rps16 and trnL). Values above branches or at nodes represent posterior probability (PP) and bootstrap percentage (BP) support. Dashes represent branches that collapsed in MP strict consensus tree. Colors represent geographic distribution of taxa; in bold are Angraecum sensu Garay species. Abbreviations in brackets denote sections sensu Garay: Aca = Acaulia, Agd = Angraecoides, Ang = Angraecum, Arc = Arachnangraecum, Bor = Boryangraecum, Chl = Chlorangraecum, Fil = Filangis, Gom = Gomphocentrum, Had = Hadrangis, Hum = Humblotiangraecum, Lem = Lemurangis, Lep = Lepervenchea, Pct = Pectinaria, Per = Perrierangraecum, Psj = Pseudojumellea.

https://doi.org/10.1371/journal.pone.0163194.g001

The phylogeny obtained from ITS2 lacked resolution and was slightly incongruent with the combined plastid phylogeny (S2 Fig). Notably, the Malagasy Angraecinae genera Aeranthes, Jumellea, and Lemurorchis were embedded within Angraecum, rendering it paraphyletic. Nonetheless, all represented clades within Angraecum appeared to be congruent with those in the plastid data (Fig 1). Our combined plastid and morphological analyses yielded a topology that is congruent with the combined plastid analyses with slight differences in branch supports (S3 Fig). Within Angraecum, clade M (PP 1.0, BP 75) gained support while clade D became weaker. The position of Lemurorchis was ambiguous, forming a polytomy with the AeranthesJumellea clade in the molecular analyses while it was embedded with Angraecum in the combined analyses. The morphological analyses alone supported the monophyly of Angraecinae sensu Carlsward et al. [26], with many polytomies observed within the clade (result not shown). According to the morphological phylogeny, most genera were monophyletic (Aeranthes, Jumellea, and most of the African genera) except Angraecum, where conflicts were also observed when compared to the plastid tree. The combined molecular (plastid and nuclear) and morphological analyses yielded a result that is incongruent with the combined plastid phylogeny, but similar to the nuclear topology (result not shown).

Divergence time estimates

The maximum credibility tree of the calibrated relaxed molecular clock analysis of Angraecinae is shown in S4 Fig. Our results suggest that Angraecinae shared a most recent common ancestor (MRCA) in the late Oligocene (~ 26.1 Ma, giving a node height highest posterior density (HPD) intervals at 95% ranging from 18.2–33.5 Ma), and started to diverge in the early Miocene (S5 Table). Diversification started at approximately 21 Ma (95% HPD: 14.6–27.8 Ma) for clade I and 17.1 Ma (95% HPD: 11.7–23.3 Ma) for clade II. In clade I, the African-American clade B diverged at approximately 19.46 Ma (95% HPD: 13.7–25.9 Ma), while the Malagasy-IOI clade A diverged at approximately 18.1 Ma (95% HPD: 12.1–24.5 Ma). Within clade II, Angraecum, Aeranthes and Jumellea started to diversify at around 15.5 Ma (95% HPD: 10.4–20.6 Ma), 10.3 Ma (95% HPD: 5.8–15.1 Ma), and 7 Ma (95% HPD: 4.7–11 Ma) respectively. By counting the number of taxa (nodes) on our calibrated maximum credibility tree (S4 Fig), it appears that 60% of the speciation events within the orchid genera Aeranthes, Angraecum and Jumellea happened during the Pliocene-Pleistocene. Angraecum section Hadrangis which is endemic to the Mascarene Islands diverged at approximately 1.66 Ma (95% HPD: 0.4–3.3 Ma), and the divergence time for the two species A. bracteosum and A. striatum that are endemic to Reunion is estimated at 0.2 Ma (95% HPD: 0–0.7 Ma).

Species diversification

Results from BiSSE showed that the second model (μ0 ~ μ1) received the best AIC score for flower color and labellum position (S6 Table). The green and white colors had similar rates of speciation (Fig 2A), while the uppermost labellum showed a higher speciation rate compared to the lowermost one (Fig 2B). Results from MuSSE showed that the third transition model (1 ←2 ←3 ↔ 4) had the best AIC score for flower size, while the full model received the best score for spur length (S6 Table); these models were used to test diversification. The diversification model with extinction rates equal between states received the best score for flower size and spur length, suggesting that character states have an effect on speciation. Large flowers showed higher speciation rates compared to small, medium and very large flowers (Fig 2C). All spur length states had a similar effect on speciation rates (Fig 2D).

thumbnail
Fig 2.

Posterior probability distributions for the speciation rates (in Ma) of morphological characters using equal rate speciation (μ0 ~ μ1) with the BiSSE model and equal rate extinction (λi ~ λj) with the MuSSE model: flower color (A), labellum position (B), flower size (C), and spur length (D). Abbreviation: v, very.

https://doi.org/10.1371/journal.pone.0163194.g002

The speciation/extinction model from BAMM revealed 5 distinct configuration shifts from the 95% credible set (Fig 3, percentages indicated at top of trees) of which 66% of the samples in the posterior distribution showed no shift, 14% showed a single shift at the node of clade II, 6.6% had one within the Aeranthes clade, 5.9% had one shift at the branch of Beclardia macrostachya, and 3.6% of the posterior distribution had one shift at the branch of clade I. Our BAMM results showed a gradual decline of speciation rate-through-time (RTT) for Angraecinae, starting from approximately 0.39 Ma during the Miocene to 0.23 Ma towards the present (Fig 3).

thumbnail
Fig 3. Configuration shifts from the 95% credible set sampled by BAMM from the Angraecinae phylogeny and evolutionary rates through time.

The intensity of colors on branches reflects the relative probability density of speciation rates (cool colors = slow, warm = fast). Black circles denote the position of the macroevolutionary regime shifts present in each sample. Blue curve indicates the mean speciation rate-through-time trajectory of Angraecinae in million years. Values above dendograms (f: configuration sampled frequency) indicate the marginal probability of rate shifts observed on branches across the posterior distribution of macroevolutionary rate shift configurations.

https://doi.org/10.1371/journal.pone.0163194.g003

The phenotypic/evolutionary model showed 32 distinct configurations on spur length. The five that received the best sample frequencies are displayed in Fig 4. Four main shifts are observed, one at the branch of the Eichlerangraecum clade, one at the branch of clade A, one at the branch of Angraecum appendiculatum, and one at the branch of Angraecum corrugatum. The RTT phenotypic evolution showed an increased rate in spur length starting from 0.02 Ma during the Pliocene to 0.08 Ma in the Pleistocene and to the present (Fig 4). No shift has been detected regarding flower size within Angraecinae and the RTT phenotypic evolution was constant (S5 Fig).

thumbnail
Fig 4. Best configuration shifts from the 95% credible set sampled by BAMM for the evolution of spur length across the phylogeny of Angraecinae.

Color intensity on branches reflects the relative probability density of the instantaneous rate of phenotypic evolution. Black circles denote the position of the macroevolutionary regime shifts present in each sample. Blue curve denotes the mean evolution rate-through-time trajectory.

https://doi.org/10.1371/journal.pone.0163194.g004

Ancestral state reconstructions

Our results showed that a labellum in the lower position is plesiomorphic in Angraecinae, while an upper labellum is apomorphic and evolved at least five times independently (Fig 5). White flowers appear to be symplesiomorphic in Angraecinae, while green flowers are apomorphic and evolved independently several times in Angraecinae and twice in Angraecum (S6A Fig). A medium flower size appears to be the ancestral state in Angraecinae, while large and small flowers are derived (S6B Fig). Long spur is the ancestral state and short spur is derived and arose several times independently (S6C Fig). Our results showed that the color (green), the size (small and large), and the spur length (short) of flowers are homoplasic within Angraecum. The taxa that showed uppermost labella appeared to be monophyletic and received very strong support in the phylogeny (Angraecoides, Dolabrifolia, Pectinariella, and Angraecum clades H to M).

thumbnail
Fig 5.

Ancestral state reconstructions of floral characters in Angraecinae implemented in ‘diversitree’; colors represent character states of the labellum position (uppermost and lowermost) and pie charts represent the probability of ancestral states at nodes. C denotes taxa illustrated in the pictures to the right to represent the flower shape of each clade except for Jumellea which is represented by Jumellea comorensis (not sampled in the phylogeny). Photo: Andriananjamanantsoa.

https://doi.org/10.1371/journal.pone.0163194.g005

Discussion

Systematics of Angraecinae

Our results confirmed the monophyly of the Angraecinae and the distinctiveness of the Malagasy–IOI clades from the African–American clade [26,28]. Our results also demonstrated the polyphyly of Angraecum s.l. Indeed, in addition to the African sections already transferred to new genera [38], four Malagasy species were embedded within clade A: Angraecum cf. humile, A. perparvulum, A. pterophyllum, and A. rhynchoglossum, belonging to sections Lemurangis, Nana, Pectinaria, and Acaulia, respectively [31]. Micheneau et al. [28] first reported the existence of a rare Angraecum species from Madagascar and the Mascarenes (A. sp. TP84, section Nana) in clade A. In our treatment, this taxon proved to be unrelated to ours and embedded within Lemurella (result not shown). To keep Angraecum monophyletic, these species must be removed from the genus.

The problem we encountered while amplifying the ITS region was primarily caused by endophytes: instead of amplifying orchid DNA, the primers amplified fungi. This was also reported by several authors working on Orchidaceae (e.g. [26]). The incongruence observed with the morphological phylogeny illustrates the existence of homoplasies in Angraecinae and the difficulty of delineating natural groups using morphology alone in this group. The variation in branch support between the combined plastid and combined plastid + morphological analyses are probably due to synapomorphic characters that were observed in some clades but were lacking in others, as well as the numerous plesiomorphic characters that are encountered in subtribe Angraecinae. The poor resolutions observed within or between clades were probably due in part to a lack of sampling or caused by a low number of markers at the positions analysed. More regions are needed to increase the resolution of the phylogeny, especially the relationships between clades.

Systematics of Angraecum

The paraphyly observed in the ITS2 data may be due to a lack of sampling and to low genetic variation. Carlsward et al. [26] had detected paralogy in the ITS region in the Malagasy Angraecinae and most notably in clade II and therefore excluded this region from subsequent analyses, including the genera Aeranthes, Jumellea, Lemurorchis, Oeoniella and Sobennikoffia. Since most of the sequences we used to reconstruct the ITS phylogeny came from their work, except for Angraecum which came mostly from our samples, and given that only ITS2 was sequenced by us, we believe that it is difficult to reconcile the data at this point. Given the potential variability of ITS sequences, it might be worth attempting to resolve the problems of paralogy and primer.

The plastid results also showed that Angraecum is paraphyletic. Two Malagasy genera, Oeoniella and Sobennikoffia, are nested within this clade, as was reported by Carlsward et al. [26]. However, Aeranthes and Jumellea form natural groups sister to Angraecum as also reported by Micheneau et al. [28]. Schlechter [33] admitted the close resemblance between Oeoniella and Angraecum, but emphasized the differences in column shape and stipe length. These two genera differ from Angraecum in their labellum shape. If we look closer at Oeoniella and unfold the labellum, the flower looks similar to Angraecum eburneum, except that the perianth is soft, the labellum spurless, and the ovary untwisted. Sobennikoffia, a small genus of four species, was previously included within Angraecum [25] until transferred to a new genus by Schlechter [32] because of the three-lobed labellum.

Of the 16 sections of Garay [31] represented in our phylogeny, only one is monophyletic, the Mascarene section Hadrangis; all other sections are paraphyletic (Gomphocentrum, Lemurangis and Lepervenchea) or polyphyletic (Acaulia, Angraecum, Angraecoides, Arachnangraecum, Boryangraecum, Filangis, Humblotiangraecum, Pectinaria, Perrierangraecum and Pseudojumellea). Micheneau et al. [28] mentioned the unnaturalness of Garay’s sections and pointed out the complexity of dealing with the sections with small greenish flowers. We decided to not consider A. nanum within clade I because of their morphological differences. Although these species produce green flowers, A. nanum, the sole member of section Nana, is characterized by tiny plant with racemose inflorescences and minute flowers, while clade I is characterized by erect or pendent plants with one-flowered inflorescences and larger flowers. The position of A. amplexicaule is ambiguous but in the BEAST analyses it was embedded within clade L (S4 Fig). Morphologically, A. amplexicaule has an inflorescence and flower shape similar to those of species in clade L, except that in clade L, the habit is more robust and the leaves are coriaceous. Despite the weak support, we decided to include A. pseudofilicornu and A. scottianum in clade J because all species in this clade share the same inflorescence type and flower shape except that the sepals and petals are reduced and the habit crassulescent for the two species. Furthermore, in the ITS2 topology A. scottianum was embedded within the clade (S2 Fig).

The morphological characters (type of inflorescence, flower size and color, and spur length) used by many authors (e.g. [31,32,34,42,44]) to delineate sections in Angraecum are of limited taxonomic interest and their positions on the phylogeny are not coherent. Spur length and flower size are generally correlated (small flowers have short spur and large flowers have long ones), with the exception of A. appendiculatum and A. corrugatum. These two Mascarene species, often considered peloric forms of A. arachnites and A. conchoglossum, respectively [28,31,68], instead appear to have lost or changed the genes responsible for labellum and spur development [69]. The phenotypic plasticity observed in these two species was considered to be the result of species radiation [28], but colonization of new habitats (with new selection regimes) might affect gene expression that is responsible for floral development [7074]. Chang et al. [72] showed that the size and shape of sepal/petal/labellum in Oncidium were regulated by the OMADS5 gene. These Angraecum species are spurless, the labellum was changed from suborbicular concave, typical in section Arachnangraecum, to linear-lanceolate, and they became self-pollinated [75]. There is no evidence here of a loss of bilateral symmetry (there is still a slight zygomorphy) that would be expected in peloric flowers. Alternately, variation in spur length could be due to genetic drift [76,77]. For instance, Stewart et al. [42] observed that spur length in subspecies of A. eburneum was shorter the further away from Madagascar a subspecies was. One of the possible causes of flower similarity could be convergent evolution as observed in clade C and the two African genera Dolabrifolia and Pectinariella previously placed in Angraecum section Pectinaria [31], or in section Humblotiangraecum (clade G) and the African genus Eichlerangraecum (Fig 1).

Labellum position: a missing character to delineate sections in Angraecum

After evaluating several morphological characters (S2 Table), we noted that the position of the labellum, uppermost or lowermost, was a good indicator of clade boundaries in Angraecum (Fig 5). The degree of resupination in an orchid flower can vary from 0° to approximately 360° depending on the inflorescence [78], and may be specific to a taxon. The 180° resupination results in the lowermost labellum position and a 360° resupination (double twist of pedicel or ovary) in the uppermost labellum position. The position of the labellum has been used in Orchidaceae to delimit sections within genus Bulbophyllum [79], and in Gesneriaceae to delimit the genus Alloplectus [80]. The newly defined African genera [38] Angraecoides, Dolabrifolia, and Pectinariella are also characterized by an uppermost labellum. Our results showed that Clades C to G have lowermost labella, while clades H to M are composed exclusively of species with uppermost labella. The dispersion of the species of section Angraecum sensu Garay (Fig 1: clade E vs clade K) in the phylogeny is a concrete example demonstrating the usefulness of this character to delimit sections. The labellum position is difficult to observe on herbarium sheets and often the only way to clearly see it is on living specimens. This could explain the fact that none of the original species descriptions mentioned this character [31,32,34]. Given these new findings, species descriptions and generic classifications should be updated.

Temporal framework and paleoclimate events in Angraecum

Our age estimate for Angraecinae and Angraecum is approximately 4 Ma older than that obtained by Micheneau et al. [81]; the age of two endemic species of the Reunion Island (A. bracteosum and A. striatum) are younger in our results, however. This can be explained by our calibration. We set the root of the tree to 35 ± 4.9 Ma following Gustafsson et al. [59], while Micheneau et al. [81] followed Ramirez et al. [82] and fixed it at 30.37 ± 3.44 Ma. The main difference between the two calibrations is based on the number of fossils used in the analyses: Ramirez et al. [82] used a single fossil, whereas Gustafsson et al. [59] included three to calibrate their phylogeny of the Orchidaceae. Furthermore, Micheneau et al. [81] used island ages as constraints in their analysis. Since no fossil was available to directly calibrate our phylogeny, these age estimates remain approximations [83]. Nonetheless, our calibration is consistent with the divergence time of the Vandeae estimated by Givnish et al. [39], and our divergence time estimate for Angraecum corresponds well with the diversification age of most Malagasy Angiosperm endemic genera [84].

During the Miocene, the climate of Madagascar shifted gradually from cool dry to warm humid. Contraction of the arid forest and expansion of the current tropical forest in Madagascar have been documented as a result of the northern migration of the island towards the equator [84,85]. This migration, together with the establishment of the trade winds, and later the monsoons [85,86], increased moisture levels throughout eastern Madagascar, thus promoting diversification of epiphytes like Angraecum and the most diverse orchid genus Bulbophyllum [79]. The high number of speciation events observed during the Pliocene-Pleistocene [5.333 Ma to 0.011 Ma] could be explained by glaciation events, where the climate fluctuated repeatedly from cold to warm [87], or the increase in the monsoon systems’ activity during the late Miocene [86]. This increase in orchid taxa which coincides with the radiation bursts observed in Malagasy tree ferns during the same period [3] may reflect the presence of humid and warm climate promoting diversification. Moreover, species richness in some Malagasy vertebrate groups were also reported to be influenced by clade age and of particular relevance, adaptability to rainforest habitats [88]. However, the 60% increase in speciation events observed within the orchid genera Aeranthes, Angraecum and Jumellea during the Pliocene-Pleistocene is difficult to reconcile with the smooth steady decreases in species diversity depicted by the BAMM analysis (Fig 3), and require further research to resolve.

Diversification in Angraecum

The BiSSE and MuSSE results revealed that flower size and labellum position appear associated with the diversification of Angraecinae and Angraecum. An uppermost labellum concurred with a higher speciation rate compared to a lowermost labellum (Fig 2B). Medium and large flowers were associated with higher speciation rates when compared to minute, small or very large flowers (Fig 2C). The overlap in the posterior probabilities observed in flower color and spur length (Fig 2A and 2D) could be interpreted as an equal rate of speciation between color and length categories. Floral divergence in Orchidaceae has been associated with pollinator shifts [89]. Fischer et al. [79] pointed out the importance of flower orientation in plant evolution and on species diversification. Lowermost labella serve as landing platforms for pollinators [79], while uppermost labella are associated with either autogamy or a switch to pollinating insects that prefer walking rather than flying [90]. Long spurs have been shown to be associated with specific pollinators in Angraecum [91,92]. Because of this specificity, long-spurred flowers are more efficiently pollinated [9396], though it does not necessarily result in a greater speciation rate, as appears to be the case in Angraecum.

The speciation/extinction analyses obtained from the BAMM model could be resumed to three best configuration shifts in Angraecinae, one shift at the MRCA of clade I, one at the MRCA of clade II, and one within the Aeranthes clade (Fig 3). BAMM detected general shifts where diversification could potentially have arisen (at the MRCA of Aeranthes, Angraecum, and Jumellea), but could not detect evidence in specific clades associated with diversification regimes that we expected within Angraecum (the shift from lowermost to uppermost labella (Fig 2B), or the high speciation rate detected by BiSSE with flower color (Fig 2A). A lack of performance of the statistical models used in BAMM was pointed out by Rabosky and Goldberg [67]. The shift detected within Aeranthes could be a bias of incomplete sampling (Fig 3). Within Jumellea we selectively sampled at least one representative of each section of the genus [41], while sampling was more random in Aeranthes. Better samples are required before any conclusion on the diversification regimes operating in this clade can be drawn.

The phenotypic/evolution analyses obtained from BAMM model showed that there was no shift associated with labellum size, while at least four shifts were detected with spur length (Fig 4). This leads us to conclude that flower size is evolutionary constant within clades in the whole subtribe Angraecinae, while the evolution of spur length is more variable in some clades than in others. Rakotoarivelo et al. [41] pointed out the lability of spur length in Jumellea, and its inefficiency on delimiting sections. Since flower size is generally conserved within clades, it is difficult to ascertain the extent to which this character may influence speciation, even though this character usually is associated with pollinator type. The shifts observed in clade C and J with spur length evolution suggest that these clades are the most phenotypically diverse in the whole Angraecinae. However, it does not necessary mean that species diversification is associated with the rate of spur length evolution. If we look at clade C, there is no phylogenetic signal showing that short or long spurs led to more speciation, which excludes all hypotheses of high rates of diversification associated with this character. The rate shifts observed in clade J were caused by the two Mascarene species A. appendiculatum, and A. corrugatum.

Our BAMM analyses showed a slowdown in time of the diversification rate in Angraecinae in general (Fig 3). To test for homogeneity in the diversification rate through time, we calculated the gamma statistic of Pybus and Harvey [15] as implemented in the R package ‘ape’. Our significantly negative result (γ = – 2.204816) rejected the null hypothesis for a constant-rate, suggesting that the speciation rate was initially higher but slowed gradually through time, which is congruent with the BAMM results. This early burst in lineages could be explained by the fact that Angraecum species developed a crassulacean acid metabolism (CAM) that allows them to tolerate desiccation [97]. Whitman et al. [98] reported that lithophytic Angraecinae are tolerant to limited moisture availability, and chronic bushfire did not kill the population of Angraecum sororium and Jumellea rigida but only reduced their expansion. But diversification slowdown could be also the result of ecological limitation, competition, access to pollinators, species carrying capacity [99,100], or maintenance of niche similarity [101]. Lineage diversification is usually high right after colonization of a new niche, but slows through time as niches get occupied and ecological conditions for speciation decrease [8]. Scantlebury et al. [22] reported that diversification slowdown in Malagasy fauna is a general pattern of adaptive radiation like that observed in amphibians and birds [5,8]. However, it has been shown that the diversification pattern of the tropical rain forest has been an accumulation of lineages through time and not sudden adaptive radiations [20]. Our results appear to support this hypothesis.

It is intriguing that each clade in Angraecum is associated with specific floral characters. Aeranthes and Jumellea are species rich and occupy the same ecological niche as Angraecum. The three genera diverged approximately at the same time, but Angraecum has more species. If we look at the morphological differentiation between the three genera, Jumellea and Aeranthes have distinctive characters that characterize them as clades, while Angraecum has variable characters specific to each subclade. The diversity of Angraecum could perhaps be partly explained by its morphological evolution. Taxonomic diversity has been demonstrated to be complemented by morphological and ecological diversity [102]. Pollinators could have played a significant role in this diversification process. Little is known about Angraecum pollinators, but we know now that it is pollinated by different kinds of pollinators [75,103105], not only hawk-moths as originally thought [91,92,106]. Micheneau et al. [28] reported that the radiation of Angraecum in the Mascarenes was caused by a change in pollinators on these islands. The absence of the original pollinators is considered to have resulted in auto-pollination in some Angraecinae species [75,107], Hermans and Hermans [108], which could be associated with a decrease or loss of rewards such as fragrance or nectar in the species [75].

Conclusion

The present study presents the most comprehensive phylogenetic reconstruction of genus Angraecum to date including all the sections sensu Garay except the African section Afrangraecum. With the African sections removed [38] Angraecum currently includes ca. 190 taxa (including varieties). Our results confirmed the paraphyly and polyphyly of Angraecum sections, but showed certain features of morphology to be consistent with phylogeny. Principally, the position of the labellum, lowermost or uppermost, allowed us to delineate several clades. An updated systematic revision of the genus is required considering these findings. Our study revealed that many characters are associated with species diversification of Angraecum, the orientation of the labellum being one. However, our analyses failed to detect shifts that could have been caused by morphological diversification. Overall, the evolution and diversification of Angraecum resulted from gradual species accumulation through time rather than rapid radiation.

Supporting Information

S1 Fig.

K-means cascade plot showing the group attributed to (A) flower size and (B) spur length for each partition. Partitions with four (4) groups were selected for analyses for both characters.

https://doi.org/10.1371/journal.pone.0163194.s001

(TIF)

S2 Fig. Phylogenetic relationships of Angraecinae, 50% Bayesian majority-rule consensus tree from the nuclear ribosomal sequences (ITS2).

Values above branches or at nodes represent posterior probability (PP) and bootstrap percentage (BP) support. Dashes represent branches that collapsed in the maximum parsimony strict consensus tree; taxa in bold are Angraecum sensu Garay. Abbreviations in brackets denote sections sensu Garay: Aca = Acaulia, Ang = Angraecum, Arc = Arachnangraecum, Bor = Boryangraecum, Fil = Filangis, Hum = Humblotiangraecum, Pct = Pectinaria, Per = Perrierangraecum, Psj = Pseudojumellea.

https://doi.org/10.1371/journal.pone.0163194.s002

(TIF)

S3 Fig. Phylogenetic relationships within subtribe Angraecinae, 50% Bayesian majority-rule consensus tree from plastid (matK, rps16 and trnL) and morphological data.

Values above branches or at nodes represent posterior probability (PP) and bootstrap percentage (BP) support. Dashes represent branches that collapsed in the maximum parsimony strict consensus tree; taxa in bold are Angraecum sensu Garay. Abbreviations in brackets denote sections sensu Garay: Aca = Acaulia, Agd = Angraecoides, Ang = Angraecum, Arc = Arachnangraecum, Bor = Boryangraecum, Chl = Chlorangraecum, Fil = Filangis, Gom = Gomphocentrum, Had = Hadrangis, Hum = Humblotiangraecum, Lem = Lemurangis, Lep = Lepervenchea, Pct = Pectinaria, Per = Perrierangraecum, Psj = Pseudojumellea.

https://doi.org/10.1371/journal.pone.0163194.s003

(TIF)

S4 Fig. Maximum credibility tree of the calibrated relaxed molecular clock analysis of Angraecinae inferred from combined plastid matK, rps16 and trnL sequences.

Posterior probabilities are displayed above branches in italics; node ages are indicated in bold, with blue bars representing the 95% highest height probability densities (HPD) of the node. Gray dashed-line indicates the Pliocene events. Abbreviations: A, Angraecum; At, Aeranthes; J, Jumellea.

https://doi.org/10.1371/journal.pone.0163194.s004

(TIF)

S5 Fig. Configuration shifts from the 95% credible set sampled by BAMM for the evolution of spur length across a phylogenetic tree of Angraecinae with evolutionary rates through time.

The intensity of colors on branches reflects the instantaneous rate of phenotypic evolution (cool colors = slow, warm = fast). The red curve illustrates the mean speciation rate-through-time trajectory of Angraecinae in million years.

https://doi.org/10.1371/journal.pone.0163194.s005

(TIF)

S6 Fig. Ancestral state reconstructions of floral characters in Angraecinae implemented in ‘diversitree’; colors represent character states and pie charts represent the probability of ancestral states at node.

Characters: (A) flower color, (B) flower size and (C) spur length.

https://doi.org/10.1371/journal.pone.0163194.s006

(TIF)

S1 Table. Voucher and sources of DNA data.

Taxa, geographical origin, voucher information, and NCBI or BOLD accession numbers (matK, rps16, trnL, ITS) for samples used in the study; sequences that are from different individuals are presented with voucher information in brackets. Blank denote unavailable sequences, “unvouchered” are sequences from living collections used in previous studies [1,3]. Unvouchered specimens have been identified by H.N. Andriananjamanantsoa and J. Andriatiana (TAN). Abbreviations: Af, Africa; Am, America; As, Asia; Com, Comoros; Md, Madagascar; Reu, Reunion; Sey, Seychelles.

https://doi.org/10.1371/journal.pone.0163194.s007

(CSV)

S2 Table. Morphological character description.

Description of vegetative and floral morphological characters in Angraecinae.

https://doi.org/10.1371/journal.pone.0163194.s008

(CSV)

S3 Table. Morphological character matrix.

Morphological character matrix used in the phylogenetic reconstruction.

https://doi.org/10.1371/journal.pone.0163194.s009

(CSV)

S4 Table. Characters used in BAMM analyses.

Continuous characters used for the phenotypic/evolution model, K-means partitioning and fractioning used for the speciation/extinction model implemented in BAMM, values for flower size and spur length were log-transformed and fraction represents the percentage of samples used in the analyses for each clade (from 0 to 1).

https://doi.org/10.1371/journal.pone.0163194.s010

(CSV)

S5 Table. Ages of crown nodes of relevant clades presented in the phylogeny of Angraecinae inferred from BEAST.

Comparison between the mean time to most recent common ancestor (MRCA) estimates and the node height highest posterior density (HPD) intervals at 95%.

https://doi.org/10.1371/journal.pone.0163194.s011

(CSV)

S6 Table. Comparaison of diversificaiton model used for the BiSSE and MuSSE analyses.

Arrow represents transition allowed between states: ↔, reversible, →, irreversible. Parameters: lambda, speciation rate; mu, extinction rate; q, character transition rate. Character and character states: color of flower (0, green; 1, white); position of labellum (0, uppermost; 1, lowermost); size of flower (1, small; 2, medium; 3, large; 4, very large); spur length (1, short; 2, medium; 3, long; 4, very long). Abbreviations: AIC, Akaike information criterion; M, model tested; lnLik, Log likelihood.

https://doi.org/10.1371/journal.pone.0163194.s012

(CSV)

Acknowledgments

The authors thank A. Schuiteman (Royal Botanic Gardens, Kew) and M. Pignal (Museum of Paris) for their help in accessing herbarium specimens. We would like to thank the University of Antananarivo and Madagascar National Parks for granting research permits. We are thankful to M. Kuzmina and the Canadian Center for DNA Barcoding for help with sequencing and the Jodrell laboratory for providing genetic material for this study. We are very thankful to C. Micheneau, R. Lei, S. Joly, J. Saarela, F.J. Lapointe, E. Gagnon, M. Light, and S. Renaut for their assistance, comments and expertise. Computations were made on the supercomputer “Cottos” and “Briaree” from the University of Montreal, managed by Calcul Québec and Compute Canada.

Author Contributions

  1. Conceptualization: HNA SE EEL LB.
  2. Data curation: HNA.
  3. Formal analysis: HNA.
  4. Funding acquisition: EEL LB.
  5. Investigation: HNA.
  6. Methodology: HNA LB.
  7. Project administration: HNA EEL LB.
  8. Resources: SE EEL LB.
  9. Supervision: EEL LB.
  10. Validation: HNA SE EEL LB.
  11. Visualization: HNA SE LB.
  12. Writing – original draft: HNA.
  13. Writing – review & editing: HNA SE EEL LB.

References

  1. 1. Goodman S, Benstead J. Updated estimates of biotic diversity and endemism for Madagascar. Oryx. 2005;39(1):73–7.
  2. 2. Martin RD. Adaptive radiation and behaviour of the Malagasy lemurs. Philos Trans—R Soc Biol Sci. 1972;264(862):295–352.
  3. 3. Janssen T, Bystriakova N, Rakotondrainibe F, Coomes D, Labat J-N, Schneider H. Neoendemism in Madagascan scaly tree ferns results from recent, coincident diversification bursts. Evolution. 2008;62(8):1876–89. pmid:18452578
  4. 4. Townsend TM, Vieites DR, Glaw F, Vences M. Testing species-level diversification hypotheses in Madagascar: the case of microendemic Brookesia leaf chameleons. Syst Biol. 2009;58(6):641–56. pmid:20525615
  5. 5. Vences M, Wollenberg K, Vieites D, Lees D. Madagascar as a model region of species diversification. Trends Ecol Evol. 2009;24(8):456–65. pmid:19500874
  6. 6. Anthony F, Diniz LC, Combes M-C, Lashermes P. Adaptive radiation in Coffea subgenus Coffea L. (Rubiaceae) in Africa and Madagascar. Plant Syst Evol. 2010;285(1–2):51–64.
  7. 7. Jønsson K, Fabre P-H, Fritz H, Etienne R, Ricklefs R, Jørgensen T, et al. Ecological and evolutionary determinants for the adaptive radiation of the Madagascan vangas. Proc Natl Acad Sci U S A. 2012;109(17):6620–5. pmid:22505736
  8. 8. Reddy S, Driskell A, Rabosky D, Hackett SJ, Schulenberg TS. Diversification and the adaptive radiation of the vangas of Madagascar. Proc—R Soc Biol Sci. 2012;279(1735):2062–71.
  9. 9. Rakotoarinivo M, Blach Overgaard A, Baker W, Dransfield J, Moat J, Svenning J-C. Palaeo-precipitation is a major determinant of palm species richness patterns across Madagascar: a tropical biodiversity hotspot. Proc—R Soc Biol Sci. 2013;280:1–9.
  10. 10. Christidis L, Goodman S, Naughton K, Appleton B, Russo D. Insights into the evolution of a cryptic radiation of bats: dispersal and ecological radiation of Malagasy Miniopterus (Chiroptera: Miniopteridae). PLoS One. 2014;9(3):e92440. pmid:24642892
  11. 11. Goldberg J, Knapp M, Emberson R, Townsend JI, Trewick S, Colgan D. Species radiation of carabid beetles (Broscini: Mecodema) in New Zealand. PLoS One. 2014;9(1):e86185. pmid:24465949
  12. 12. Goodman KR, Kelley JP, Welter SC, Roderick GK, Elias DO. Rapid diversification of sexual signals in Hawaiian Nesosydne planthoppers (Hemiptera: Delphacidae): the relative role of neutral and selective forces. J Evol Biol. 2015;28(2):415–27. pmid:25535672
  13. 13. Lewis D, Sperling FAH, Nakahara S, Cotton A, Kawahara A, Condamine F. Role of Caribbean Islands in the diversification and biogeography of neotropical heraclides swallowtails. Cladistics. 2015;31(3):291–314.
  14. 14. Pagel M. Detecting correlated evolution on phylogenies: a general method for the comparative analysis of discrete characters. Proc—R Soc Biol Sci. 1994;255(1342):37–45.
  15. 15. Pybus OG, Harvey PH. Testing macro-evolutionary models using incomplete molecular phylogenies. Proc—R Soc Biol Sci. 2000;267(1459):2267–72.
  16. 16. Maddison WP, Midford P, Otto SP. Estimating a binary character’s effect on speciation and extinction. Syst Biol. 2007;56(5):701–10. pmid:17849325
  17. 17. FitzJohn R, Maddison WP, Otto SP. Estimating trait-dependent speciation and extinction rates from incompletely resolved phylogenies. Syst Biol. 2009;58(6):595–611. pmid:20525612
  18. 18. Rabosky D, Donnellan S, Grundler M, Lovette I. Analysis and visualization of complex macroevolutionary dynamics: an example from Australian scincid lizards. Syst Biol. 2014;63(4):610–27. pmid:24682412
  19. 19. Schluter D. The ecology of adaptive radiation. New York: Oxford University Press Inc.; 2000.
  20. 20. Couvreur TLP, Forest F, Baker W. Origin and global diversification patterns of tropical rain forests: inferences from a complete genus-level phylogeny of palms. BMC Biol. 2011;9:44. pmid:21679405
  21. 21. Rabosky D, Santini F, Eastman J, Smith S, Sidlauskas B, Chang J, et al. Rates of speciation and morphological evolution are correlated across the largest vertebrate radiation. Nat Commun. 2013;4:1958. pmid:23739623
  22. 22. Scantlebury D. Diversification rates have declined in the Malagasy herpetofauna. Proc—R Soc Biol Sci. 2013;280(1766):20131109.
  23. 23. Myers N, Mittermeier R, Mittermeier C, da Fonseca GAB, Kent J. Biodiversity hotspots for conservation priorities. Nature. 2000;403(6772):853–8. pmid:10706275
  24. 24. Chase MW, Cameron KM, Freudenstein JV, Pridgeon AM, Salazar G, van den Berg C. An updated classification of Orchidaceae. Bot J Linn Soc. 2015;177(2):151–74.
  25. 25. Govaerts R, Campacci M, Baptista D, Cribb P, George A, Krenuz K, et al. World checklist of Orchidaceae. Facilitated by the Royal Botanic Garden, Kew Published on the internet. Available: https://doi.org/http://appskeworg/wcsp/ Accessed May 2015.
  26. 26. Carlsward B, Stern W, Bytebier B. Comparative vegetative anatomy and systematics of the angraecoids (Vandeae, Orchidaceae) with an emphasis on the leafless habit. Bot J Linn Soc. 2006;151(2):165–218.
  27. 27. Dressler RL. Phylogeny and classification of the orchid family. Cambridge University Press; 1993.
  28. 28. Micheneau C, Carlsward B, Fay M, Bytebier B, Pailler T, Chase M. Phylogenetics and biogeography of Mascarene angraecoid orchids (Vandeae, Orchidaceae). Mol Phylogenet Evol. 2008;46(3):908–22. pmid:18272406
  29. 29. Cribb P. Erasanthe (Orchidaceae, Epidendroideae, Vandeae, Aerangidinae), a new endemic orchid genus from Madagascar. Adansonia. 2007;29(1):27–30.
  30. 30. Cribb P, Carlsward B. New combinations in Aerangis, Diaphananthe and Podangis (Orchidaceae, subtribe Angraecinae). Phytotaxa. 2012;71(1):42–7.
  31. 31. Garay LA. Systematics of the genus Angraecum (Orchidaceae). Kew Bull. 1973;28:495–516.
  32. 32. Schlechter R. Orchidaceae Perrierianae. Beih Zum Bot Cent. 1925;33:1–391.
  33. 33. Schlechter R. Versuch einer naturlichen Neuordnung der afrikanischen agraekoiden Orchidaceen. Beih Zum Bot Cent. 1918;36:62–181.
  34. 34. Perrier de la Bâthie H. 49e Famille-Orchidées. Humbert H. Tananarive: Imprimerie Officielle; 1941. 202–321 p. (Flore de Madagascar).
  35. 35. Summerhayes VS. African orchids: XXVI. Kew Bull. 1958;13(2):260–77.
  36. 36. Simo-Droissart M, Micheneau C, Sonké B, Droissart V, Plunkett GM, Lowry PP, et al. Morphometrics and molecular phylogenetics of the continental African species of Angraecum section Pectinaria (Orchidaceae). Plant Ecol Evol. 2013;146(3):295–309.
  37. 37. Szlachetko D, Romowicz A. Dolabrifolia, un nouveau genre d’orchidees de l’alliance Angraecum. Richardiana. 2007;7(2):53–4.
  38. 38. Szlachetko D, Tukałło P, Mytnik Ejsmont J, Grochocka E. Reclassification of the Angraecum-alliance (Orchidaceae, Vandoideae) based on molecular and morphological data. Biodivers Res Conserv. 2013;29(1):1–23.
  39. 39. Givnish T, Spalink D, Ames M, Lyon S, Hunter S, Zuluaga A, et al. Orchid phylogenomics and multiple drivers of their extraordinary diversification. Proc—R Soc Biol Sci. 2015;282(1814):20151553.
  40. 40. Chase MW. Silica gel: an ideal material for field preservation of leaf samples for DNA studies. Taxon. 1991;40(2):215–20.
  41. 41. Rakotoarivelo FP, Razafimandimbison SG, Mallet B, Faliniaina L, Pailler T. Molecular systematics and evolutionary trends and relationships in the genus Jumellea (Orchidaceae): implications for its species limits. Taxon. 2012;61(3):534–44.
  42. 42. Stewart J, Hermans J, Campbell B. Angraecoid orchids: species from the African region. Portland: Timber Press Inc.; 2006.
  43. 43. Du Puy DJ, Moat J. A refined classification of the primary vegetation of Madagascar based on the underlying geology: using GIS to map its distribution and to assess its conservation status. In: Biogéographie de Madagascar. ORSTOM. Paris; 1996. p. 205–18.
  44. 44. Cribb P, Hermans J. Field guide to the orchids of Madagascar. Kew: Royal Botanic Gardens; 2009.
  45. 45. Doyle JJ. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem Bull. 1987;19:11–5.
  46. 46. Yu J, Xue J-H, Zhou S-L. New universal matK primers for DNA barcoding angiosperms. J Syst Evol. 2011;49(3):176–81.
  47. 47. Oxelman B, Lidn M, Berglund D. Chloroplast rps16 intron phylogeny of the tribe Sileneae (Caryophyllaceae). Plant Syst Evol. 1997;206(1–4):393–410.
  48. 48. Taberlet P, Gielly L, Pautou G, Bouvet J. Universal primers for amplification of three non-coding regions of chloroplast DNA. Plant Mol Biol. 1991;17(5):1105–9. pmid:1932684
  49. 49. Chen S, Yao H, Han J, Liu C, Song J, Shi L, et al. Validation of the ITS2 region as a novel DNA barcode for identifying medicinal plant species. PLoS One. 2010;5(1):e8613. pmid:20062805
  50. 50. Silva WAJ, Costa MC, Valente V, Sousa JF, Paco-Larson ML, Espreafico EM, et al. PCR template preparation for capillary DNA sequencing. BioTechniques. 2001;30(3):537–42. pmid:11256382
  51. 51. Gouy M, Guindon S, Gascuel O. SeaView version 4: a multiplatform graphical user interface for sequence alignment and phylogenetic tree building. Mol Biol Evol. 2010;27(2):221–4. pmid:19854763
  52. 52. Hall TA. BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic acids symposium series. 1999.
  53. 53. Swofford DL. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods), version 4.0 Beta. Sunderland Mass Sinauer. 2002.
  54. 54. Ronquist F, Huelsenbeck J. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19(2):1572–4.
  55. 55. Posada D, Crandall K. Modeltest: testing the model of DNA substitution. Bioinformatics. 1998;14:817–8. pmid:9918953
  56. 56. Akaike HA. A new look at the statistical model identification. IEEE Trans Autom Control. 1974;19(6):716–23.
  57. 57. Drummond A, Rambaut A. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol Biol. 2007;7:214. pmid:17996036
  58. 58. Iles WJD, Smith S, Gandolfo M, Graham S. Monocot fossils suitable for molecular dating analyses. Bot J Linn Soc. 2015;178(3):346–74.
  59. 59. Gustafsson ALS, Verola C, Antonelli A. Reassessing the temporal evolution of orchids with new fossils and a Bayesian relaxed clock, with implications for the diversification of the rare South American genus Hoffmannseggella (Orchidaceae: Epidendroideae). BMC Evol Biol. 2010;10(1):177.
  60. 60. Freudenstein JV, Chase MW. Phylogenetic relationships in Epidendroideae (Orchidaceae), one of the great flowering plant radiations: progressive specialization and diversification. Ann Bot. 2015;115:665–81. pmid:25578379
  61. 61. FitzJohn R. Diversitree: comparative phylogenetic analyses of diversification in R. Methods Ecol Evol. 2012;3(6):1084–92.
  62. 62. Rabosky D, Kolokotronis S-O. Automatic detection of key innovations, rate shifts, and diversity-dependence on phylogenetic trees. PLoS One. 2014;9(2):e89543. pmid:24586858
  63. 63. Heibl C. PHYLOCH: R language tree plotting tools and interfaces to diverse phylogenetic software packages. 2013.
  64. 64. Legendre P, Legendre L. Numerical ecology. 3rd English edition. Elsevier, Amsterdam; 2012.
  65. 65. Borcard D, Gillet F, Legendre P. Numerical ecology with R. Springer, New York; 2011.
  66. 66. Maddison W, FitzJohn R. The unsolved challenge to phylogenetic correlation tests for categorical characters. Syst Biol. 2015;64(1):127–36. pmid:25209222
  67. 67. Rabosky D, Goldberg E. Model inadequacy and mistaken inferences of trait-dependent speciation. Syst Biol. 2015;64(2):340–55. pmid:25601943
  68. 68. Hermans J, Cribb P. The Angraecum germinyanum complex. Orchid Rev. 2005;113:90–7.
  69. 69. Hossain I, Levy C, Rindge C. Investigating KNOX gene expression in Aquilegia petal spur development. Jouranl Emerg Investig. 2014;1–6.
  70. 70. Theiβen G, Saedler H. MADS-box genes in plant ontogeny and phylogeny: Haeckel’s “biogenetic law” revisited. Curr Opin Genet Dev. 1995;5(5):628–39. pmid:8664551
  71. 71. Hsu HF, Huang CH, Chou LT, Yang CH. Ectopic expression of an orchid (Oncidium Gower Ramsey) AGL6-like gene promotes flowering by activating flowering time genes in Arabidopsis thaliana. Plant Cell Physiol. 2003;44(8):783–94. pmid:12941870
  72. 72. Chang Y-Y, Kao N-H, Li J-Y, Hsu W-H, Liang Y-L, Wu J-W. Characterization of the possible roles for B class MADS box genes in regulation of perianth formation in orchid. Plant Physiol. 2010;152(2):837–53. pmid:20018605
  73. 73. Pan ZJ, Cheng CC, Tsai WC, Chung MC, Chen WH, Hu JM, et al. The duplicated B-class MADS-box genes display dualistic characters in orchid floral organ identity and growth. Plant Cell Physiol. 2011;52(9):1515–31. pmid:21757456
  74. 74. Ding L, Wang Y, Yu H. Overexpression of DOSOC1, an ortholog of Arabidopsis SOC1, promotes flowering in the orchid Dendrobium Chao Parya Smile. Plant Cell Physiol. 2013;54(4):595–608. pmid:23396600
  75. 75. Pailler T, Tournebize R, Henze F. Guide des orchidées de la Réunion: 52 espèces indigènes décrites. Université de la Réunion; 2013.
  76. 76. Mallet B, Martos F, Blambert L, Pailler T, Humeau L, Canestrelli D. Evidence for isolation-by-habitat among populations of an epiphytic orchid species on a small Oceanic Island. PLoS One. 2014;9(2):e87469. pmid:24498329
  77. 77. Luo Y, Widmer A, Karrenberg S. The roles of genetic drift and natural selection in quantitative trait divergence along an altitudinal gradient in Arabidopsis thaliana. Heredity. 2015;114(2):220–8. pmid:25293874
  78. 78. Arditti J. Resupination. Lankesteriana Int J Orchid. 2003;7:95–6.
  79. 79. Fischer G, Gravendeel B, Sieder A, Andriantiana J, Heiselmayer P, Cribb P, et al. Evolution of resupination in Malagasy species of Bulbophyllum (Orchidaceae). Mol Phylogenet Evol. 2007;45(1):358–76. pmid:17716924
  80. 80. Clark JL. A preliminary phylogeny of Alloplectus (Gesneriaceae): implications for the evolution of flower resupination. Syst Bot. 2003;28(2):365–75.
  81. 81. Micheneau C, Fournel J, Warren B, Hugel S, Gauvin Bialecki A, Pailler T, et al. Orthoptera, a new order of pollinator. Ann Bot. 2010;105(3):355–64. pmid:20067913
  82. 82. Ramírez S, Gravendeel B, Singer R, Marshall C, Pierce N. Dating the origin of the Orchidaceae from a fossil orchid with its pollinator. Nature. 2007;448(7157):1042–5. pmid:17728756
  83. 83. Graur D, Martin W. Reading the entrails of chickens: molecular timescales of evolution and the illusion of precision. Trends Genet. 2004;20(2):80–6. pmid:14746989
  84. 84. Buerki S, Devey D, Callmander M, Phillipson P, Forest F. Spatio-temporal history of the endemic genera of Madagascar. Bot J Linn Soc. 2013;171(2):304–29.
  85. 85. Samonds KE, Godfrey LR, Ali JR, Goodman SM, Vences M, Sutherland MR, et al. Spatial and temoral arrival patterns of Madagascar’s vertebrate fauna explained by distance, ocena currents, and ancestor type. PNAS. 2012;109(14):5352–7. pmid:22431643
  86. 86. Samonds KE, Godfrey LR, Ali JR, Goodman SM, Vences M, Sutherland MR, et al. Imperfect isolation: factors and filters shaping Madagascar’s extant vertebrate fauna. PLoS One. 2013;8(4):e62086. pmid:23626770
  87. 87. Ruddiman WF. Earth’s climate: past and future. Macmillan; 2001.
  88. 88. Crottini A, Madsen O, Poux C, StrauB A, Vieites DR, Vences M. Vertebrate time-tree elucidates the biogeographic pattern of a major biotic around the K-T boundary in Madagascar. PNAS. 2012;109(14):5358–63. pmid:22431616
  89. 89. Peter C, Johnson S. A pollinator shift explains floral divergence in an orchid species complex in South Africa. Ann Bot. 2014;113(2):277–88. pmid:24107684
  90. 90. Rudall P, Bateman R. Roles of synorganisation, zygomorphy and heterotopy in floral evolution: the gynostemium and labellum of orchids and other lilioid monocots. Biol Rev. 2002;77(3):403–41. pmid:12227521
  91. 91. Nilsson LA, Jonsson L, Ralison L, Randrianjohany E. Angraecoid orchids and hawkmoths in central Madagascar: specialized pollination systems and generalist foragers. Biotropica. 1987;19:310–8.
  92. 92. Wasserthal L. The pollinators of the Malagasy star orchids Angraecum sesquipedale, A. sororium and A. compactum and the evolution of extremely long spurs by pollinator shift. Bot Acta. 1997;110(5):343–59.
  93. 93. Fulton M, Hodges SA. Floral isolation between Aquilegia formosa and Aquilegia pubescens. Proc—R Soc Biol Sci. 1999;266(1435):2247–52.
  94. 94. Hodges S, Fulton M, Yang J, Whittall J. Verne grant and evolutionary studies of Aquilegia. New Phytol. 2004;161(1):113–20.
  95. 95. Whittall J, Hodges S. Pollinator shifts drive increasingly long nectar spurs in columbine flowers. Nature. 2007;447(7145):706–19. pmid:17554306
  96. 96. Boberg E, Alexandersson R, Jonsson M, Maad J, Ågren J, Nilsson LA. Pollinator shifts and the evolution of spur length in the moth-pollinated orchid Platanthera bifolia. Ann Bot. 2014;113(2):267–75. pmid:24169591
  97. 97. Kluge M, Vinson B, Ziegler H. Ecophysiological studies on orchids of Madagascar: incidence and plasticity of crassulacean acid metabolism in species of the genus Angraecum Bory. Plant Ecol. 1998;135(1):43–57.
  98. 98. Whitman M, Medler M, Randriamanindry JJ, Rabakonadrianina E. Conservation of Madagascar’s granite outcrop orchids: the influence of fire and moisture. Lankesteriana Int J Orchid. 2011;11(1):55–67.
  99. 99. Gavrilets S, Vose A. Dynamic patterns of adaptive radiation. Proc Natl Acad Sci U S A. 2005;102(50):18040–5. pmid:16330783
  100. 100. Phillimore A, Price T, Barton N. Density-dependent cladogenesis in birds. PLoS Biol. 2008;6(3):e71. pmid:18366256
  101. 101. Moen D, Morlon H. Why does diversification slow down? Trends Ecol Evol. 2014;29(4):190–7. pmid:24612774
  102. 102. Shi J, Rabosky D. Speciation dynamics during the global radiation of extant bats. Evolution. 2015;69(6):1528–45. pmid:25958922
  103. 103. Micheneau C, Fournel J, Pailler T. Bird pollination in an angraecoid orchid on Reunion Island (Mascarene Archipelago, Indian Ocean). Ann Bot. 2006;97(6):965–74. pmid:16554310
  104. 104. Micheneau C, Fournel J, Humeau L, Pailler T. Orchid-bird interactions: a case study from Angraecum (Vandeae, Angraecinae) and Zosterops (white-eyes, Zosteropidae) on Reunion Island. Botany. 2008;86(10):1143–51.
  105. 105. Micheneau C, Johnson S, Fay M. Orchid pollination: from Darwin to the present day. Bot J Linn Soc. 2009;161(1):1–19.
  106. 106. Nilsson LA, Jonsson L, Rason L, Randrianjohany E. Monophily and pollination mechanisms in Angraecum arachnites Schltr.(Orchidaceae) in a guild of long‐tongued hawk‐moths (Sphingidae) in Madagasca. Biol J Linn Soc. 1985;26(1):1–19.
  107. 107. Micheneau C, Fournel J, Gauvin Bialecki A, Pailler T. Auto-pollination in a long-spurred endemic orchid (Jumellea stenophylla) on Reunion Island (Mascarene Archipelago, Indian Ocean). Plant Syst Evol. 2008;272(1–4):11–22.
  108. 108. Hermans C, Hermans J. In search of the star of Africa. Orchids Bull AOS. 2013;82(9):544–50.