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

A Set of 100 Chloroplast DNA Primer Pairs to Study Population Genetics and Phylogeny in Monocotyledons

  • Nora Scarcelli ,

    nora.scarcelli@ird.fr

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

  • Adeline Barnaud,

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

  • Wolf Eiserhardt,

    Affiliation Ecoinformatics and Biodiversity Group, Department of Biological Science, Aarhus University, Aarhus, Denmark

  • Urs A. Treier,

    Affiliation Ecoinformatics and Biodiversity Group, Department of Biological Science, Aarhus University, Aarhus, Denmark

  • Marie Seveno,

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

  • Amélie d'Anfray,

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

  • Yves Vigouroux,

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

  • Jean-Christophe Pintaud

    Affiliation UMR Diversité, Adaptation et Développement des Plantes (DIADE), Institut de Recherche pour le Développement (IRD), Montpellier, France

Abstract

Chloroplast DNA sequences are of great interest for population genetics and phylogenetic studies. However, only a small set of markers are commonly used. Most of them have been designed for amplification in a large range of Angiosperms and are located in the Large Single Copy (LSC). Here we developed a new set of 100 primer pairs optimized for amplification in Monocotyledons. Primer pairs amplify coding (exon) and non-coding regions (intron and intergenic spacer). They span the different chloroplast regions: 72 are located in the LSC, 13 in the Small Single Copy (SSC) and 15 in the Inverted Repeat region (IR). Amplification and sequencing were tested in 13 species of Monocotyledons: Dioscorea abyssinica, D. praehensilis, D. rotundata, D. dumetorum, D. bulbifera, Trichopus sempervirens (Dioscoreaceae), Phoenix canariensis, P. dactylifera, Astrocaryum scopatum, A. murumuru, Ceroxylon echinulatum (Arecaceae), Digitaria excilis and Pennisetum glaucum (Poaceae). The diversity found in Dioscorea, Digitaria and Pennisetum mainly corresponded to Single Nucleotide Polymorphism (SNP) while the diversity found in Arecaceae also comprises Variable Number Tandem Repeat (VNTR). We observed that the most variable loci (rps15-ycf1, rpl32-ccsA, ndhF-rpl32, ndhG-ndhI and ccsA) are located in the SSC. Through the analysis of the genetic structure of a wild-cultivated species complex in Dioscorea, we demonstrated that this new set of primers is of great interest for population genetics and we anticipate that it will also be useful for phylogeny and bar-coding studies.

Introduction

The knowledge of the chloroplast genome structure and sequence variation in Monocotyledons is still partial and unbalanced. There are currently 25 completely sequenced chloroplast genomes of Monocotyledons available in GenBank [1] but 17 of them are of Poales, and important orders like Liliales, Commelinales and Zingiberales lack complete chloroplast sequences. Comparative genomic analyses of the chloroplast DNA (cpDNA) relevant to Monocotyledons are scarce [2], [3], [4] and mostly focused on grasses and allied groups [5], [6], [7], [8], [9], [10], [11]. A few monocotyledonous species are documented for many genes [12] while numerous species are documented for a few genes only and non-coding regions, including the most commonly used markers for phylogenetic inference and genetic bar-coding like rbcL [13], atpB [14], trnL-F [15], matK, psbA-trnH, rpoC1, rpoB-trnC, psbK-psbI, atpF-atpH, atpH-atpI [16], [17], [18], [19], [20], [21]. Most other regions of the Large Single Copy (LSC) have been investigated in particular taxa, for example clpP intron2 in Yucca [22], rps2 in Tiphonium [23], petN-psbM in Elaeocharis [24], psbD-trnT in Arum [25], psbB and psbC in Vanilla [26], psbZ-trnfM in Livistona [27], or accD in Hexalectris [28], to mention a few studies. Variation within the slowly evolving Inverted Repeat region (IR) [29] has received little attention in Monocotyledons [30], [31], a large part of it being represented only by the complete chloroplast sequences. Within the Small Single Copy (SSC), there is limited information outside the extensively used ndhF gene [32], with only few studies using ycf1, rpl32-trnL and ndhA [33], [34], [35], [36].

Moreover, available sets of primers for direct sequencing of chloroplast regions in Angiosperms mostly focus on non-coding regions of the LSC [37], [38], [39] while published information on primers for genes is very dispersed [40].

The possibility of screening a large number of loci is useful to detect polymorphic Small Inversions, microsatellites and minisatellites (i.e. Variable Number Tandem Repeats, VNTR) in species complexes and at the population level [1], [37]. Minute and medium size inversions are frequent features of the non-coding cpDNA [41], [42], detectable only through sequencing and showing intraspecific variability [1], [43]. Microsatellites are also widespread structures in non-coding cpDNA that became important population genetics markers [44]. The most common and most widely used microsatellites are mononucleotide repeats [45]. Longer motifs, in particular minisatellites, are comparatively rare, but also proved to be valuable markers [46], [47], [48].

Here we propose a large set of primer pairs optimized for PCR amplification and overlapping sequencing in Monocotyledons. Primers pairs are distributed throughout the whole chloroplast genome and include exons, introns and Intergenic Spacers (IGS) with contrasted mutation rates and evolutionary patterns. They are thus suitable for a wide range of studies from higher-level phylogeny to population genetics. As an example, we used the newly defined primer pairs to study intra-specific cpDNA diversity of three different yam species (Dioscorea spp.)

Materials and Methods

Primer definition

The complete sequence of six Monocotyledons chloroplast genomes were downloaded from GenBank, namely Acorus calamus (NC_007407), Dioscorea elephantipes (NC_009601), Lemna minor (NC_010109), Oryza nivara (NC_005973), Phalaenopsis aphrodite (NC_007499) and Zea mays (NC_001666).

Segments of these sequences equivalent to two to six genes were aligned using the program Geneious [49]. Consensus primers anchored in exons were designed using Primer3 [50] incorporated in Geneious, in order to amplify IGS, introns or exons. A total of 105 primers pairs were designed, and 100 successfully amplified: 72 in the Large Single Copy region (LSC), 13 in the Small Single Copy (SSC) and 15 in the Inverted Repeat region (IR). Primer sequences, annealing temperature for PCR amplification, and amplification results are summarized in Table S1.

Test for amplification

Amplification was tested in 13 species of Monocotyledons: 6 Dioscoreaceae species (1 individual each of Dioscorea abyssinica, D. praehensilis, D. rotundata, D. dumetorum, D. bulbifera and Trichopus sempervirens), 5 Arecaceae species (1 individual each of Phoenix canariensis, P. dactylifera, Astrocaryum scopatum, A. murumuru and 2 individuals of Ceroxylon echinulatum), Digitaria excilis (5 individuals) and Pennisetum glaucum (6 individuals). Sequences have been deposited in GenBank under accession number JF705257-JF705858, JF745569-JF745769 and JF758190-JF758233.

Amplification was done according to the recommended protocols using either GoTaq (Promega) in its buffer with 5 mM of dNTPs for D. excilis and P. glaucum or Failsafe enzyme mix (Epicentre) in premix E for Dioscoreaceae and Arecaceae species. Reaction was done in 25 µL with 25 ng of DNA. The initial denaturation (94°C, 3 min) was followed by 35 cycles of denaturation (94°C, 30 s), annealing (Tm, 30 s) and elongation (72°C, 1 min) and by a final elongation step (72°C, 10 min). Amplification was checked on agarose gel.

Sequencing

PCR products were purified using Ampure (Agencourt) following the recommended protocol. The sequencing PCRs were done using the BigDye terminator kit (Applied Biosystems). PCR products were purified using CleanSeq (Agencourt) and were run on ABI prism 3130 (Applied Biosystems). Note that for D. excilis and P. glaucum, only a subset of the PCR products was sequenced. Dioscoreaceae and Arecaceae species were sequenced in forward and reverse direction while D. excilis and P. glaucum were sequenced in forward direction only.

Data analysis

Sequences were aligned with the program Geneious [49]. Intra-generic diversity was estimated within species or between closely-related species as the number of SNP and the number of Variable Number Tandem Repeats (VNTR). The number of SNP was standardised to 1 kb but length variable parts (e.g., gaps or VNTR) were subtracted from the total length of the alignment. For D. excilis and P. glaucum, intra-generic diversity was estimated within the analysed species. For Dioscoreaceae, it was estimated between closely-related species (D. rotundata, D. abyssinica and D. praehensilis of subgenus Eniantophyllum). For Arecaceae it was estimated as the mean of the diversity found within each species pair in Phoenix, Astrocaryum and Ceroxylon.

Due to high inter-generic divergence in Dioscoreaceae [51] causing alignment difficulties in non-coding regions with T. sempervirens, nucleotide diversity was only estimated between two distant species of Dioscorea, D. abyssinica and D. elephantipes. For Arecaceae, an average of three inter-generic comparison, between Phoenix (subfamily Coryphoideae) and Astrocaryum (subfamilly Arecoideae), Phoenix and Ceroxylon (subfamily Ceroxyloideae) and Ceroxylon and Astrocaryum was calculated. Within Poaceae, inter-generic diversity was not estimated for D. excilis and P. glaucum because only a part of the loci were sequenced. Instead, inter-generic diversity was estimated between Oryza sativa and Zea mays using the GenBank sequences but restricted to those parts theoretically amplified by the primer pairs tested in the present study.

Comparisons of genetic diversity between SSC, LSC and IR, and between introns, exons and IGS were performed with Kruskal-Wallis tests using the R environment [52], function kruskal.test.

Example of use for population genetic analysis

We analysed the genetic structure of three yam species (Dioscorea spp.) forming a crop-wild relatives complex in Western Africa. The main cultivated yam species in West-Africa is D. rotundata. In this region, yam is a staple food but is also culturally extremely important [53]. The wild relatives of D. rotundata are D. abyssinica and D. praehensilis [54], [55]. The three species are genetically different but can hybridize [56].

One sample of each species has been previously sequenced (see above). Based on these sequences, 19 polymorphic loci were identified showing a total of 21 SNP. These 19 loci have been tested on eight additional individuals (four D. abyssinica and four D. praehensilis) to selected those loci for which polymorphisms were specific to either D. abyssinica or D. praehensilis; namely ccsA-Exon, ccsA-ndhD, ndhH-Exon, psbD-Exon and rrn4,5-trnN.

Finally, a total of 160 Dioscorea samples have been amplified using the selected five primers pairs. The sampling included 66 D. abyssinica, 39 D. praehensilis and 55 D. rotundata collected in Benin. A list of individuals and sampling locations is given in the supplementary data file (Table S2). Sequences have been deposited in GenBank under accession number JF757240-JF758189. The five loci revealed six SNP (two for rrn4,5-trnN and one each for the other loci). A chlorotype is defined as a combination of SNP located on the chloroplast, i.e. a haplotype based on chloroplast SNP. Here, the combinations of the six SNP revealed five chlorotypes. The repartition of chlorotype frequencies among species was compared with a chi-squared test. A MSN, Minimum Spanning Network [57], with chlorotypes was constructed using Haplophyle [58]. MSN illustrates the evolutionary relationships between chlorotypes as a network where the branches represent the differences between sequences data.

Results and Discussion

Development of new chloroplast primers

Of the 105 primer pairs designed to sequence the chloroplast genome, 100 amplified consistently and produced good quality sequences. Primers were designed to amplify a wide range of monocotyledons species and we tested them on various species of different genera (D. abyssinica, D. praehensilis, D. rotundata, D. dumetorum, D. bulbifera, T. sempervirens, P. canariensis, P. dactylifera, A. scopatum, A. murumuru, C. echinulatum, D. excilis, P. glaucum). Amplification success was 85% (Table S1) which was very similar to the expected mean amplification of 88% derived from the sequences deposited in GenBank used to design the primers (95% for A. calamus, 95% for D. elephantipes, 97% for L. minor, 80% for O. nivara, 88% for P. aphrodite and 80% for Z. mays). Indeed, due to structural changes (inversions, gene loss, etc.) some primers pairs are expected not to amplify in some species. For example, because of the loss of ycf2 and accD in O. nivara and Z. mays, we do not expect amplification with primers pairs rpl23-ycf2, ycf2-ndhB, accD-psaI, rbcL-accD and accD-Exon on these two species.

Primers amplified coding regions (exon 20%), non-coding regions (IGS 35%, intron 9%) and mixed regions (exon+intron 10%, IGS+genes 25%). 75% of these regions were located in the LSC, 12% in the SSC and 15% in the IR.

Sequence diversity

We obtained a total of 1174 kb sequence data. The analysis covered 78 kb of the chloroplast genome for Dioscoreaceae (51% of the D. elephantipes cpDNA), 70 kb for Arecaceae (44% of the P. dactylifera cpDNA), 34 kb for Digitaria (25% of the O. nivara cpDNA) and 20 kb for Pennisetum (15% of the O. nivara cpDNA).

A summary of intra- and inter-generic diversity results are presented in Table 1. Detailed results are given in supplementary data file (Table S3).

Intra-generic diversity

We found on average a SNP each 1700 bp within the three Arecaceae genera, each 2800 bp between the three Dioscorea species D. abyssinica, D. praehensilis and D. rotundata, each 8900 bp among the six P. glaucum samples and each 9600 bp among the five D. excilis samples. These very low levels of intra-generic diversity in the studied Poaceae suggest a strong bottleneck effects in such cultivated populations. There were few polymorphic microsatellites in Dioscorea, D. excilis and P. glaucum, all mononucleotide, while the Arecaceae exhibit a high number of mono-, di- and 4–8-nucleotide microsatellites as well as minisatellites (Table 2). A total of 66 VNTR were found in palms, 77% of them located in IGS, 23% in intron and none in exon (Table 3). The 51 polymorphic mononucleotide microsatellites encountered within genera and species of palms can be compared with the 342 homopolymers of 7 bp or longer found in the complete chloroplast genome of Phoenix dactylfiera [1].

thumbnail
Table 2. Comparison of the number and motifs of polymorphic VNTR observed at the intra-generic level.

https://doi.org/10.1371/journal.pone.0019954.t002

thumbnail
Table 3. Polynucleotide VNTRs with repeat number >2 and polymorphic within genera in palms.

https://doi.org/10.1371/journal.pone.0019954.t003

Interestingly, Arecaceae species exhibit a much higher number of VNTR than Dioscorea species. Similar levels of mono- and dinucleotide microsatellites in Dioscorea as in closely related palm species could only be found if two distant species (D. abyssinica and D. elephantipes) were compared (data not shown). This result suggests different evolutionary histories with higher mutation rates and/or larger effective population sizes in Arecaceae than in Dioscorea species.

Inter-generic diversity

Between O. sativa and Z. mays we found a SNP each 21 bp, between D. abyssinica and D. elephantipes each 75 bp and for the three inter-generic comparisons in Arecaceae on average each 113 bp. Since Oryza sativa and Zea mays diverged about 52 MY ago [59] and the compared palm subfamilies diverged about 68–98 MY ago [60], our result confirmed a 5–6 fold faster substitution rate for cpDNA in Poaceae than in Arecaceae [61]. The genus concept in Dioscoreaceae is very different from that of Poaceae and Arecaceae. Levels of divergence between two distant species of Dioscorea was in the range of the inter-generic differentiation in Poaceae and Arecaceae, while different Dioscoreaceae genera, namely Dioscorea and Trichopus are so divergent that they are not even alignable for some IGS.

Interestingly, we did not find significant differences in number of SNP in introns vs. exons and in introns vs. IGS, neither for Dioscorea, nor for Arecaceae and Poaceae (p>0.05 for Dioscorea, Arecaceae and Poaceae). We observed a significantly higher number of SNP in IGS vs. exons only for Dioscorea (p<0.05) and in Poaceae (p<0.01). This finding highlights the very peculiar dynamics of SNP in the chloroplast genome. It can be compared with the result of Yang et al. [1] who identified 62 out of 78 SNP within the cultivar ‘Khalass’ of the date palm occurring in exons, with an unusual synonymous/non synonymous ratio of 0.94. They suggested a lack of purifying selection within heterogeneous intra-individual chloroplast populations as a possible explanation (Yang et al. 2010b).

The occurrence of SNP among the three regions of the chloroplast (LSC, SSC and IR) varies (Figure 1). LSC and SSC exhibit similar levels of diversity while IR exhibits significantly lower numbers of SNP. The difference in number of SNP is significant for LSC vs. IR and SSC vs. IR (p<0.001 for Dioscorea, Arecaceae and Poaceae) but is not significant for LSC vs. SSC (p>0.05 for Dioscorea, Arecaceae and Poaceae). Variation in SNP number in the SSC region is, however, mostly driven by the ndhF-rpl32 locus. This locus exhibits a very high genetic diversity: 92 and 118 SNP per 1 kb for Dioscorea and Poaceae, compared to the mean of 16, 10 and 62 SNP per 1 kb for Dioscorea, Arecaceae and Poaceae, respectively and of 11, 7 and 30 SNP per 1 kb in the whole chloroplast.

thumbnail
Figure 1. Inter-generic diversity found in Dioscorea, Arecaceae and Poaceae.

Inter-generic diversity was estimated as the number of SNP in the LSC, SSC and IR. Numbers of SNP have been standardised to 1 kb. Bars represent the 95% confidence intervals.

https://doi.org/10.1371/journal.pone.0019954.g001

Most published primer pairs focus on non-coding regions of the LSC [37], [38], [39]. This region is commonly used for phylogeny and bar-coding [13], [14], [15], [16], [17], [18], [19], [20]. In the present study, we observed some of the most variable loci in the SSC, namely rps15-ycf1, rpl32-ccsA and ndhF-rpl32 for Dioscorea and rps15-ycf1, ndhG-ndhI and ccsA for Arecaceae.

Polynucleotide VNTR in palms

Polynucleotide VNTR are apparently rare in Monocotyledons. They are virtually absent in Dioscorea and Poaceae, although a 22 bp minisatellite located in the trnD-trnT region, with 1–3 repeats, has been reported in Elymus [47]. A complex evolution of minisatellites was also detected in an orchid, Anacamptis, within the trnL intron [62].

Palms are outstanding for the frequency of such structures in the chloroplast genome. In this study, 12 VNTR were recorded in the genus Astrocaryum, with motif length varying from 2 to 26 bp. There was, however, considerable variation in VNTRs abundance among genera of palms (Table 3). In Phoenix, only two polynucleotide VNTRs were detected, namely 2 minisatellites of 12 and 20 bp. Within C. echinulatum, there was no polymorphism at the level of the polynucleotide VNTR, and only 9 of the 51 mononucleotide microsatellites were polymorphic. We note, however, that only two individuals have been compared and VNTRs occurrence might be higher. Differences between Astrocaryum and Ceroxylon might be explained by differences in divergence time between the pairs of individuals compared (less than 2 MY in Ceroxylon, about 7 MY in Astrocaryum) and also by a higher sequence variability in Astrocaryum and other Bactridinae compared with Ceroxyleae and Phoeniceae [63], [64], [65], [66].

Thus, polynucleotide VNTRs have a great potential in palms for population genetic studies and species delimitation. They have already been used with success in several studies. For example, the dodecanucleotide minisatellite of the psbZ-trnfM locus showed fixed private haplotypes that allowed the separation of closely related Phoenix species and tracking interspecific hybridization [48]. The tetranucleotide microsatellite of the trnQ-rps16 locus allowed tracing seed flow between the wild and cultivated compartments of the peach palm (Bactris gasipaes) in western Ecuador and proved to be much more informative than a mononucleotide microsatellite present in the same locus [67].

We note, however, that among the 13 polynucleotide VNTRs with motifs longer than 3 bp found in the palm sampling (Table 2), five are direct repeats, i.e. incipient VNTR with 1–2 units of the motif. Direct repeats are common in non-coding cpDNA [68], and are probably the starting point of more repeated polynucleotide VNTR, although few loci undergo this evolution. For example, a sequence of 8 bp in the rbcL-accD spacer was found unrepeated in Phoenix and Ceroxylon but showed 2–3 tandem repeats in Astrocaryum (Table 3). Some minisatellites also originate from inversions [62].

As already noted above, the comparison of a limited number of individuals per family, as in the present study, might considerably underestimates the actual number of VNTR in a given taxa. Indeed, an alignment of the 174 palm sequences deposited in GenBank of the locus trnQ-rps16 alone (1.1 kb) revealed 16 intra-generic direct repeat polymorphisms 5–22 bp long, a mononucleotide microsatellite with 8–17 repeats, a dinucleotide microsatellite with 3–6 repeats, a tetranucleotide microsatellite with 2–6 repeats and a 26 bp minisatellite with 1–4 repeats. The last structure is polymorphic in a single group, the subtribe Linospadicinae, restricted to the south-west Pacific [69].

For detailed studies of VNTR variation in a particular group, it is therefore advisable to begin with the sequencing of a significant number of samples, in order to evaluate accurately the existing polymorphism in the target locus.

Example of use for population genetic analysis

CpDNA is generally inherited by only one parent (usually the mother in angiosperms). It is haploid and it generally lacks recombination [70]. CpDNA is therefore of great interest for population genetics studies, including parentage analysis, hybridization, population structure and phylogeography [44].

Here we used the new primer set to study the genetic structure of a yam crop-wild relatives' complex (cultivated: D. rotundata, wild: D. abyssinica and D. praehensilis) in Benin, Western Africa. After screening more than half of the chloroplast genome the set of informative loci retained to study this species complex included four (out of five) loci from the SSC and IR regions, among which three were exons. This emphasizes again how interesting the rarely studied SSC and IR regions are and confirms that exons are not less variable than introns or IGS, as far as it concerns SNP.

We found five chlorotypes among the 160 sequenced Dioscorea individuals that showed significantly different frequencies among the three species (Fig. 2a, p<0.001 for all pairwise comparisons, chi-squared tests). Chlorotypes 2 and 3 were specific to D. abyssinica; chlorotypes 4 and 5 are specific to D. praehensilis; while the most common chlorotype 1 was found in all three species. Chlorotypes 1, 2 and 3 as well as chlorotypes 4 and 5 were closely related with only one SNP separating them (Fig. 2b).

thumbnail
Figure 2. Chlorotypes observed within Dioscorea species.

(A) Distribution of observed chlorotypes and (B) minimum spanning network (MSN) representing the relationship among chlorotypes. In (B), the size of the circle is proportional to the chlorotype occurrence and each line represents a SNP between the two connected chlorotypes. Each SNP has been labelled with the name of the primer pairs needed for its amplification and the corresponding base change.

https://doi.org/10.1371/journal.pone.0019954.g002

The cultivated species D. rotundata harboured only chlorotype 1 and thus was less diverse than its wild relatives D. abyssinica and D. praehensilis. However, because chlorotype 1 was shared by all three species, we cannot conclude on the maternal origin of D. rotundata.

Our results showed that SNP revealed by sequencing can successfully be used to study the diversity of the crop-wild relatives' complex of Dioscorea. Furthermore, the genetic diversity revealed by sequencing with five primer pairs was more informative than the genetic diversity observed using five universal chloroplast mononucleotide microsatellites [71].

We thus showed that the new primer set can reveal diversity even when microsatellites might not show polymorphism, as it was the case in the Dioscorea species complex studied. We anticipate that the use of sequencing and SNP genotyping for population genetic analysis will be even more interesting for species or species complexes showing higher genetic diversity, as in some groups of Arecaceae like Bactridinae.

Conclusion

In this paper, we present a large set of newly developed chloroplast DNA primer pairs. Compared to the previously published primer pairs [37], [38], [39], [40], this new set covers a wider range of the chloroplast genome (e.g. up to 51% of the Dioscorea cpDNA) and has been designed to optimally amplify in Monocotyledons. This new set of primer pairs spans the Large Single Copy as well as the Small Single Copy and the Inverted Repeats, and has been designed to amplify both coding (exon) and non-coding (intron, intergenic spacer) regions. This new set could be of great interest for phylogeny and bar-coding studies but also for population genetics studies.

Supporting Information

Table S1.

Primer sequences and amplification range. Primers were designed using genes alignment of Dioscorea elephantipes, Zea mays, Oryza nivara, Lemna minor, Acorus calamus and Phalaenopsis aphrodite. Amplifications were tested on different species of Dioscoreaceae, Digitaria, Pennisetum and Arecaceae.

https://doi.org/10.1371/journal.pone.0019954.s001

(DOC)

Table S2.

List of Dioscorea individuals used to test the use of the new primers pairs for population genetics studies. Table includes the chlorplotype (1 to 5) corresponding to each sample.

https://doi.org/10.1371/journal.pone.0019954.s002

(DOC)

Table S3.

Observed Intra- and inter-generic diversity. Intra-generic diversity was estimated between closely-related species for Dioscorea and Arecaceae or within species (Digitaria excilis and Pennisetum glaucum). Inter-generic diversity was estimated between different genera (Arecaceae and Poaceae) or distant species (Dioscorea). The number of SNP was standardised to 1 kb.

https://doi.org/10.1371/journal.pone.0019954.s003

(DOC)

Acknowledgments

We greatly acknowledge S. De Mita for help in GenBank submission. We thank the editor and an anonymous reviewer for their useful comments.

Author Contributions

Conceived and designed the experiments: NS AB YV JCP. Performed the experiments: NS AB MS AdA. Analyzed the data: NS AB WE UAT. Contributed reagents/materials/analysis tools: NS AB YV JCP. Wrote the paper: NS AB YV JCP.

References

  1. 1. Yang M, Zhang X, Liu G, Yin Y, Chen K, et al. (2010) The complete chloroplast genome sequence of date palm (Phoenix dactylifera L.). PLoS ONE 5: e12762.
  2. 2. Hansen D, Dastidar S, Cai Z, Penaflor C, Kuehl J, et al. (2007) Phylogenetic and evolutionary implications of complete chloroplast genome sequences of four early-diverging angiosperms: Buxus (Buxaceae), Chloranthus (Chloranthaceae), Dioscorea (Dioscoreaceae), and Illicium (Schisandraceae). Molecular Phylogenetics and Evolution 45: 547–563.
  3. 3. Mardanov A, Ravin N, Kuznetsov B, Samigullin T, Antonov A, et al. (2008) Complete sequence of the duckweed (Lemna minor) chloroplast genome: structural organization and phylogenetic relationships to othe Angiosperms. Journal of Molecular Evolution 66: 555–564.
  4. 4. Chang C, Lin H, Lin I, Chow T, Chen H, et al. (2006) The chloroplast genome of Phalaenopsis aphrodite (Orchidaceae): Comparative analysis of evolutionary rate with that of grasses and its phylogenetic implications. Molecular Biology and Evolution 23: 279–291.
  5. 5. Matsuoka Y, Yamazaki Y, Ogihara Y, Tsunewaki K (2002) Whole chloroplast genome comparison of rice, maize, and wheat: implications for chloroplast gene diversification and phylogeny of cereals. Molecular Biology and Evolution 19: 2084–2091.
  6. 6. Saski C, Lee S, Fjellheim S, Guda C, Jansen R, et al. (2007) Complete chloroplast genome sequences of Hordeum vulgare, Sorghum bicolor and Agrostis stolonifera, and comparative analyses with other grass genomes. Theoretical and applied genetics 115: 571–590.
  7. 7. Leseberg C, Duvall M (2009) The complete chloroplast genome of Coix lacryma-jobi and a comparative molecular evolutionary analysis of plastomes in cereals. Journal of Molecular Evolution 69: 311–318.
  8. 8. Guisinger M, Chumley T, Kuehl J, Boore J, Jansen R (2010) Implications of the plastid genome sequence of Typha (Typhaceae, Poales) for understanding genome evolution in Poaceae. Journal of Molecular Evolution 70: 149–166.
  9. 9. Morris L, Duvall M (2010) The chloroplast genome of Anomochloa marantoidea (Anomochlooideae; Poaceae) comprises a mixture of grass-like and unique features. American Journal of Botany 97: 620–627.
  10. 10. Bortiri E, Coleman-Derr D, Lazo G, Anderson O, Gu Y (2008) The complete chloroplast genome sequence of Brachypodium distachyon: sequence comparison and phylogenetic analysis of eight grass plastomes. BMC Research Notes 1: 61.
  11. 11. Cahoon A, Sharpe R, Mysayphonh C, Thompson E, Ward A, et al. (2010) The complete chloroplast genome of tall fescue (Lolium arundinaceum; Poaceae) and comparison of whole plastomes from the family Poaceae. American Journal of botany 97: 49–58.
  12. 12. Jansen R, Cai Z, Raubeson L, Daniell H, Depamphilis C, et al. (2007) Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proceedings of the National Academy of Sciences of the United States of America 104: 19369–19374.
  13. 13. Duvall M, Clegg M, Chase M, Clark W, Kress W, et al. (1993) Phylogenetic hypothesis for the Monocotyledons constructed from rbcL sequence data. Annals of the Missouri Botanical Garden 80: 607–619.
  14. 14. Soltis D, Soltis P, Chase M, Mort M, Albach D, et al. (2000) Angiosperm phylogeny inferred from 18S rDNA, rbcL, and atpB sequences. Botanical Journal of the Linnean Society 133: 381–461.
  15. 15. Taberlet P, Gielly L, Pautou G, Bouvet J (1991) Universal primers for amplification of three non-coding regions of chloroplast DNA. Plant Molecular Biology 17: 1105–1109.
  16. 16. Tamura M, Yamashita J, Fuse S, Haraguchi M (2004) Molecular phylogeny of monocotyledons inferred from combined analysis of plastid matK and rbcL gene sequences. Journal of Plant Research 117: 109–120.
  17. 17. Kress J, Wurdack K, Zimmer E, Weigt L, Janzen D (2005) Use of DNA barcodes to identify flowering plants. Proceedings of the National Academy of Sciences of the United States of America 102: 8369–8374.
  18. 18. Chase M, Cowan R, Hollingsworth P, van den Berg C, Madriñan S, et al. (2007) A proposal for a standardised protocol to barcode all land plants. Taxon 56: 295–299.
  19. 19. Zeng C, Zhang Y, Triplett J, Yang J, Li D (2010) Large multi-locus plastid phylogeny of the tribe Arundinarieae (Poaceae: Bambusoideae) reveals ten major lineages and low rate of molecular divergence. Molecular Phylogenetics and Evolution 56: 821–839.
  20. 20. Hollingsworth P, Forrest L, Spouge J, Hajibabaei M, Ratnasingham S, et al. (2009) A DNA barcode for land plants. Proceedings of the National Academy of Sciences of the United States of America 106:
  21. 21. Seberg O, Petersen G (2009) How many loci does it take to DNA barcode a Crocus? PLoS ONE e4598.
  22. 22. Smith C, Pellmyr O, Althoff D, Balcazar-Lara M, Leebens-Mack J, et al. (2008) Pattern and timing of diversification in Yucca (Agavaceae): specialized pollination does not escalate rates of diversification. Proceeding of the Royal Society London Series B 275: 249–258.
  23. 23. Ohi-Toma T, Wu S, Yadav S, Murata H, Murata J (2010) Molecular phylogeny of Typhonium sensu lato and its allied genera in the tribe Areae of the subfamily Aroideae (Araceae) based on sequences of six chloroplast regions. Systematic Botany 35: 244–251.
  24. 24. Hinchliff C, Roalson E (2009) Stem architecture in Eleocharis subgenus Limnochloa (Cyperaceae): evidence of dynamic morphological evolution in a group of pantropical sedges. American Journal of Botany 96: 1487–1499.
  25. 25. Espindola A, Buerki S, Bedalov M, Kupfer P, Alvarez N (2010) New insights into the phylogenetics and biogeography of Arum (Araceae): unravelling its evolutionary history. Botanical Journal of the Linnean Society 163: 14–32.
  26. 26. Cameron K, Molina M (2006) Photosystem II gene sequences of psbB and psbC clarify the phylogenetic position of Vanilla (Vanilloideae, Orchidaceae). Cladistics 22: 239–248.
  27. 27. Crisp M, Isagi Y, Kato Y, Cook L, Bowman D (2009) Livistona palms in Australia: ancient relics or opportunistic immigrants? Molecular Phylogenetics and Evolution 54: 512–523.
  28. 28. Kennedy A, Watson L (2010) Species delimitations and phylogenetic relationships within the fully myco-heterotrophic Hexalectris (Orchidaceae). Systematic Botany 35: 64–76.
  29. 29. Goremykin V, Bobrova V, Pahnke J, Troitsky A, Antonov A, et al. (1996) Noncoding sequences from the slowly evolving chloroplast Inverted Repeat in addition to rbcL data do not support Gnetalean affinities of Angiosperms. Molecular Biology and Evolution 13: 383–396.
  30. 30. Graham S, Zgurski J, McPherson M, Cherniawsky D, Saarela J, et al. (2006) Robust inference of Monocots deep phylogeny using an expanded multigene plastid data set. Aliso 22: 3–21.
  31. 31. Wang R, Cheng C, Chang C, Wu C, Su T, et al. (2008) Dynamics and evolution of the inverted repeat-large single copy junctions in the chloroplast genomes of monocots. BMC Evolutionary Biology 8: 36.
  32. 32. Olmstead R, Sweere J, Wolfe K (1993) Ninety extra nucleotide in ndhF gene of tobacco chloroplast DNA: a summary of revisions to the 1986 genome. Plant Molecular Biology 22: 1191–1193.
  33. 33. Yang J-B, Yang H-Q, L iD-Z, Wong K-M, Yang Y-M (2010) Phylogeny of Bambusa and its allies (Poaceae: Bambusoideae)inferred from nuclear GBSSI gene and plastid psbA-trnH, rpl32-trnL and rps16 intron DNA sequences. Taxon 59: 1102–1110.
  34. 34. Chase M, Williams N, de Faria A, Neubig K, Amaral Mdo C, et al. (2009) Floral convergence in Oncidiinae (Cymbidieae; Orchidaceae): an expanded concept of Gomesa and a new genus Nohawilliamsia. Annals of Botany 104: 387–402.
  35. 35. Peterson P, Romaschenko K, Johnson G (2010) A classification of the Chloridoideae (Poaceae) based on multi-gene phylogenetic trees. Molecular Phylogenetics and Evolution 55: 580–598.
  36. 36. Barrett C, Freudenstein J (2009) Patterns of Morphological and Plastid DNA Variation in the Corallorhiza striata Species Complex (Orchidaceae). Systematic Botany 34: 496–504.
  37. 37. Ebert D, Peakall R (2009) A new set of universal de novo sequencing primers for extensive coverage of noncoding chloroplast DNA: new opportunities for phylogenetic studies and cpSSR discovery. Molecular Ecology Resources 9: 777–1075.
  38. 38. Shaw J, Lickey E, Beck J, Farmer S, Liu W, et al. (2005) The tortoise and the hare II: relatively utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. American Journal of Botany 92: 142–166.
  39. 39. Shaw J, Lickey E, Schilling E, Small R (2007) Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in Angiosperms: The tortoise and the hare III. American Journal of Botany 94: 275–288.
  40. 40. Heinze B (2007) A database of PCR primers for the chloroplast genomes of higher plants. Plant Methods 3: 4.
  41. 41. Kelchner S, Wendel J (1996) Hairpins create minute inversions in non-coding regions of chloroplast DNA. Current Genetics 30: 259–262.
  42. 42. Kim K, Lee H-L (2005) Widespread occurrence of small inversions in the chloroplast genomes of land plants. Molecules and Cells 16: 104–113.
  43. 43. Whitlock B, Hale A, Groff P (2010) Intraspecific inversions pose a challenge for the trnH-psbA plant DNA barcode. PLoS ONE 5: e11533.
  44. 44. Provan J, Powell W, Hollingsworth P (2001) Chloroplast microsatellites: new tools for studies in plant ecology and evolution. Trends in Ecology and Evolution 16: 142–147.
  45. 45. Provan J, Biss P, McMeel D, Mathews S (2004) Universal primers for the amplification of chloroplast microsatellites in grasses (Poaceae). Molecular Ecology Notes 4: 262–264.
  46. 46. Cozzolino S, Cafasso D, Pellegrino G, Musacchio A, Widmer A (2003) Fine-scale phylogeographical analysis of Mediterranean Anacamptis palustris (Orchidaceae) populations based on chloroplast minisatellite and microsatellite variation. Molecular Ecology 12: 2783–2792.
  47. 47. Sun G, Ma X (2009) Nucleotide diversity and minisatellite in chloroplast Asp(GUC)–Thr(GGU) region in Elymus trachycaulus complex, Elymus alaskanus and Elymus caninus. Biochemical Systematics and Ecology 37: 67–75.
  48. 48. Pintaud J-C, Zehdi S, Couvreur T, Barrow S, Henderson S, et al. (2010) Species delimitation in the genus Phoenix (Arecaceae) based on SSR markers, with emphasis on the identity of the Date Palm (Phoenix dactylifera L.). In: Seberg O, Petersen G, Barfod A, Davis J, editors. Diversity, phylogeny, and evolution in the Monocotyledons. Denmark: Aarhus University Press. pp. 267–286.
  49. 49. Drummond A, Ashton B, Buxton S, Cheung M, Cooper A, et al. (2010) Geneious v4.7.6 Available: http://www.geneious.com. Accessed 2011 April 26.
  50. 50. Rozen S, Skaletsky H (2000) Primer3 on the WWW for general users and for biologist programmers. In: Krawetz S, Misener S, editors. Bioinformatics methods and protocols: Methods in molecular biology. Totowa: Humana Press.
  51. 51. Caddick LR, Rudall PJ, Wilkin P, Hedderson T, Chase M (2002) Phylogenetics of Dioscoreales based on combined analyses of morphological and molecular data. Botanical Journal of the Linnean Society 138: 123–144.
  52. 52. R Development Core Team (2010) R: A language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing.
  53. 53. Coursey DG (1967) Social and cultural importance. Yams, an account of the nature, origins, cultivation and utilisation of the useful members of the Dioscoreaceae. London, UK: Longmans. pp. 197–205.
  54. 54. Hamon P (1987) Structure, origine génétique des ignames cultivées du complexe Dioscorea cayenensis-rotundata et domestication des ignames en Afrique de l'Ouest. Thèse de doctorat, Université Paris XI, France.
  55. 55. Terauchi R, Chikaleke V, Thottappilly G, Hahn S (1992) Origin and phylogeny of Guinea yams as revealed by RFLP analysis of chloroplast DNA and nuclear ribosomal DNA. Theoretical and Applied Genetics 83: 743–751.
  56. 56. Scarcelli N, Tostain S, Vigouroux Y, Agbangla C, Daïnou O, et al. (2006) Farmers' use of wild relative and sexual reproduction in a vegetatively propagated crop. The case of yam in Benin. Molecular Ecology 15: 2421–2431.
  57. 57. Excoffier L, Smouse PE (1994) Using allele frequencies and geographic subdivision to reconstruct gene trees within a species: Molecular variance parsimony. Genetics 136: 343–359.
  58. 58. Sarah G, Ruiz M, Perrier X, Billot C (2007) Available: http://haplophyle.cirad.fr/index.jsp. Accessed 2011 April 26.
  59. 59. Vicentini A, Barber JC, Alisioni SS, Giussani LM, Kellog EA (2008) The age of the grasses and clusters of origins of C4 photosynthesis. Global Change Biology 14: 2963–2977.
  60. 60. Roncal J, Borchsenius F, Asmussen-Lange CB, Balslev H (2010) Divergence times in the tribe Geonomateae (Arecaceae) coincide with Tertiary geological events. In: Seberg O, Petersen G, Barfod A, Davis J, editors. pp. 245–264. Diversity, phylogeny, and evolution in the Monocotyledons: Aarhus University Press.
  61. 61. Gaut BS, Morton BR, MacCaig BC, Clegg MT (1996) Substitution rate comparisons between grasses and palms: synonymous rate differences at the nuclear gene Adh parallel rate differences at the plastid gene rbcL. Proceedings of the National Academy of Sciences of the United States of America 93: 10274–10279.
  62. 62. Cozzolino S, Cafasso D, Pellegrino G, Musacchio A, Wildmer A (2003) Molecular evolution of a plastid tandem repeat locus in an orchid lineage. Journal of Molecular Evolution 57: S41–S49.
  63. 63. Asmussen CB, Dransfield J, Deickmann V, Barfod AS, Pintaud J-C, et al. (2006) A new subfamily classification of the palm family (Arecaceae): evidence from plastid DNA phylogeny. Botanical Journal of the Linnean Society 151: 15–38.
  64. 64. Couvreur TLP, Hahn WJ, de Granville J-J, Pham J-L, Ludena B, et al. (2007) Phylogenetic relationships of the cultivated Neotropical palm Bactris gasipaes (Arecaceae) with its wild relatives inferred from chloroplast and nuclear DNA polymorphisms. Systematic Botany 32: 519–530.
  65. 65. Eiserhardt W, Pintaud J, Asmussen-Lange C, Hahn W, Bernal R, et al. (In press) Phylogeny and divergence times of Bactridinae (Arecaceae, Palmae) based on plastid and nuclear DNA sequences. Taxon.
  66. 66. Trénel P, Gustafsson MHG, Baker WJ, Asmussen-Lange CB, Dransfield J, et al. (2007) Mid-Tertiary dispersal, not Gondwanan vicariance explains distribution patterns in the wax palm subfamily (Ceroxyloideae: Arecaceae). Molecular Phylogenetics and Evolution 45: 272–288.
  67. 67. Pintaud J-C, Couvreur TLP, Lara C, Ludeña B, Pham J-L (2008) Reciprocal introgression between wild and cultivated peach palm (Bactris gasipaes Kunth, Arecaceae) in western Ecuador. In: Maxted N, Ford-Lloyd BV, Kell SP, Iriondo JM, Dulloo ME, et al., editors. Crop wild relatives conservation and use. UK: CAB International. pp. 296–308.
  68. 68. Kelchner S (2000) The evolution of non-coding chloroplast DNA and its implication in plant systematics. Annals of the Missouri Botanical Garden 87: 482–498.
  69. 69. Savolainen V, Anstett M-C, Lexer C, Hutton I, Clarkson JJ, et al. (2006) Sympatric speciation in palms on an oceanic island. Nature 441: 210–213.
  70. 70. Birky CW Jr (1995) Uniparental inheritance of mitochondrial and chloroplast genes: Mechanisms and evolution. Proceedings of the National Academy of Sciences of the United States of America 92: 11331–11338.
  71. 71. Chaïr H, Perrier X, Agbangla C, Marchand JL, Dainou O, et al. (2005) Use of cpSSRs for the characterisation of yam phylogeny in Benin. Genome 48: 674–684.