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

What does mitogenomics tell us about the evolutionary history of the Drosophila buzzatii cluster (repleta group)?

  • Nicolás Nahuel Moreyra ,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    nmoreyra@ege.fcen.uba.ar (NNM); ehasson@ege.fcen.uba.ar (EH)

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Ecología, Genética y Evolución de Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas, Ciudad Autónoma de Buenos Aires, Argentina

  • Julián Mensch,

    Roles Conceptualization, Data curation, Formal analysis, Methodology, Supervision, Validation, Writing – review & editing

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Ecología, Genética y Evolución de Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas, Ciudad Autónoma de Buenos Aires, Argentina

  • Juan Hurtado,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Writing – review & editing

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Ecología, Genética y Evolución de Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas, Ciudad Autónoma de Buenos Aires, Argentina

  • Francisca Almeida,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Supervision, Writing – review & editing

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Ecología, Genética y Evolución de Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas, Ciudad Autónoma de Buenos Aires, Argentina

  • Cecilia Laprida,

    Roles Conceptualization, Investigation, Supervision, Visualization, Writing – review & editing

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Estudios Andinos, CONICET/UBA, Ciudad Autónoma de Buenos Aires, Argentina

  • Esteban Hasson

    Roles Conceptualization, Funding acquisition, Investigation, Project administration, Resources, Supervision, Writing – original draft, Writing – review & editing

    nmoreyra@ege.fcen.uba.ar (NNM); ehasson@ege.fcen.uba.ar (EH)

    Affiliations Departamento de Ecología, Genética y Evolución, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Ciudad Autónoma de Buenos Aires, Argentina, Instituto de Ecología, Genética y Evolución de Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas, Ciudad Autónoma de Buenos Aires, Argentina

Abstract

The Drosophila repleta group is an array of more than 100 species endemic to the “New World”, many of which are cactophilic. The ability to utilize decaying cactus tissues as breeding and feeding sites is a key aspect that allowed the successful diversification of the repleta group in American deserts and arid lands. Within this group, the Drosophila buzzatii cluster is a South American clade of seven closely related species in different stages of divergence, making them a valuable model system for evolutionary research. Substantial effort has been devoted to elucidating the phylogenetic relationships among members of the D. buzzatii cluster, including molecular phylogenetic studies that have generated ambiguous results where different tree topologies have resulted dependent on the kinds of molecular marker used. Even though mitochondrial DNA regions have become useful markers in evolutionary biology and population genetics, none of the more than twenty Drosophila mitogenomes assembled so far includes this cluster. Here, we report the assembly of six complete mitogenomes of five species: D. antonietae, D. borborema, D. buzzatii, two strains of D. koepferae and D. seriema, with the aim of revisiting phylogenetic relationships and divergence times by means of mitogenomic analyses. Our recovered topology using complete mitogenomes supports the hypothesis of monophyly of the D. buzzatii cluster and shows two main clades, one including D. buzzatii and D. koepferae (both strains), and the other containing the remaining species. These results are in agreement with previous reports based on a few mitochondrial and/or nuclear genes, but conflict with the results of a recent large-scale nuclear phylogeny, indicating that nuclear and mitochondrial genomes depict different evolutionary histories.

Introduction

Almost all mitochondrial genomes, the “mitogenome”, can be assembled directly from genome or even transcriptome sequencing datasets [1, 2]. The exponential development of next-generation sequencing (NGS) technologies, together with efficient bioinformatic tools for the analysis of genomic information, has allowed efficient assembly of mitochondrial genomes, giving rise to the emergence of the mitogenomics era [3]. Mitogenomics has been very useful in illuminating phylogenetic relationships at various depths of the Tree of Life, e.g., among early branching of metazoan phyla [4], among crocodilians and their survival at the Cretaceous-Tertiary boundary [5], primates [6], the largest clade of freshwater actynopterigian fishes [7] and Anura, the largest living Amphibian group [8]. Also, mitogenomic approaches have been used to investigate evolutionary relationships in groups of closely related species (e.g. [9]). In animals, the mitochondrial genome has been a popular choice in phylogenetic and phylogeographic studies because of its mode of inheritance, rapid evolution and the fact that it does not recombine [10]. Such physical linkage implies that all regions of mitogenomes are expected to produce the same phylogeny. However, the use of different mitogenome regions or even the complete mitogenome may lead to incongruent results [11], suggesting that mitogenomics sometimes may not reflect the true species history but rather the mitochondrial history [1216]. Moreover, there is evidence suggesting that mtDNA genes are not strictly neutral markers, casting doubts on its use to infer the past history of taxa [17]. Inconsistencies across markers may result from inaccurate reconstructions or from actual differences between genes and species trees. In fact, most methods do not take into consideration that different genomic regions may have different evolutionary histories, mainly due to the occurrence of incomplete lineage sorting and introgressive hybridization [1820].

Over the last century, the Drosophila genus has been extensively studied because of the well-known advantages that several species offer as experimental models. A remarkable feature of this genus that comprises more than two thousand species [21] is its diverse ecology: some species use fruits as breeding sites, others use flowers, mushrooms, sap fluxes, and fermenting cacti (reviewed in [2225]). The adoption of decaying cacti as breeding sites occurred more than once in the evolutionary history of Drosophilidae [26, 27] and is considered a key innovation in the diversification and invasion of American deserts and arid lands by species of the Drosophila repleta group (hereafter the repleta group) [26]. Many species in this group are capable of developing in necrotic cactus tissues and feeding on cactophilic yeasts associated to the decaying process [2835].

The repleta group comprises more than one hundred species [23, 3639], however, only one of the more than twenty complete (or nearly complete) Drosophila mitogenomes assembled so far belongs to a species in this group (checked in GenBank, March 28, 2019), D. mojavensis (GenBank: BK006339.1). The latter, the first cactophilic fly to have a sequenced nuclear genome [40], is a member of the D. mulleri complex, an assemblage of species that belongs to the D. mulleri subgroup, one of the six species subgroups of the repleta group [37].

The D. buzzatii complex is the sister group of the D. mulleri complex [26]. It diversified in the Caribbean islands and South America, giving rise to the D. buzzatii cluster (hereafter the buzzatii cluster), and the D. martensis and D. stalkeri clusters [41]. The former is an ensemble of seven closely related species including D. antonietae [42], D. borborema [43], D. buzzatii [44], D. gouveai [42], D. koepferae [45], D. serido [43], and D. seriema [46]. All species are endemic to South America (Fig 1), except the semi-cosmopolitan D. buzzatii that reached a wide distribution following human mediated dispersion of prickly pears in the genus Opuntia (Caryophillales, Cactaceae) in historical times [35, 47, 48]. These species inhabit open areas of sub-Amazonian semidesert and desert regions of South America, where flies use necrotic cactus tissues as obligatory feeding and breeding resources [35, 49]. Regarding host plant use, D. buzzatii is an Opuntia specialist [31], considered an ancestral condition [26]. However, D. buzzatii has also been reared from necrotic columnar cacti [35]. The remaining species are mainly columnar dwellers although D. antonietae and D. serido can also use O. monacantha [49], while D. koepferae has also been recovered from decaying cladodes of O. monacantha, O. quimilo and O. sulphurea [31].

thumbnail
Fig 1. Geographical distribution of buzzatii cluster species modified from reference [50].

https://doi.org/10.1371/journal.pone.0220676.g001

Species of the buzzatii cluster are almost indistinguishable in external morphology, however, differences in the morphology of the male intromittent organ (aedeagus) and polytene chromosome inversions provide clues to species identification (reviewed in [35, 48, 51]). The cluster has been divided into two groups based on aedeagus morphology, the first includes D. buzzatii and the remaining species compose the so-called Drosophila serido sibling set -serido sibling set from hereafter- [48]. In turn, analysis of polytene chromosomes revealed four informative paracentric inversions that define four main lineages: inversion 5g fixed in D. buzzatii, 2j9 in D. koepferae, 2x7 shared by D. antonietae and D. serido, and 2e8 shared by D. borborema, D. gouveai, and D. seriema [41, 52]. However, neither genital morphology nor chromosomal inversions are useful for inferring basal relationships within the cluster.

Pre-genomic phylogenetic studies based on a few molecular markers generated debate since different tree topologies were recovered depending on the molecular marker used. On one hand, the mitochondrial cytochrome oxidase I (COI) and the X-linked period gene supported the hypothesis of two main clades, one including D. buzzatii and D. koepferae and another comprising the remaining species [48, 53, 54]. On the other hand, trees based on a few nuclear and mitochondrial markers supported the hypothesis that D. koepferae was sister to the serido sibling set [26, 55]. To further complicate this issue, not all the same species were analyzed in these studies. In this vein, a recent genomic level study using a large transcriptomic dataset supports the placement of D. koepferae in the serido sibling set and D. buzzatii as sister to this set [50] similar to the results of [26]. However, phylogenetic relationships within the serido sibling set could not be ascertained despite the magnitude of the dataset employed by Hurtado and co-workers [50]. Thus, our aim is to shed light on the evolutionary relationships within the buzzatii cluster by means of a mitogenomic approach.

In this paper, we report the assembly of the complete mitogenomes of D. antonietae, D. borborema, D. buzzatii, D. seriema and two strains of D. koepferae, together with the corresponding gene annotations. Unfortunately, D. gouveai and D. serido, that inhabit Brazilian arid lands, could not be included because they are difficult to obtain and are not available in Drosophila repositories. We also present a mitogenomic analysis that defines a different picture of the relationships within the buzzatii cluster with respect to the results generated with nuclear genomic data. Finally, we discuss possible causes of the discordance between nuclear and mitochondrial datasets.

Material and methods

Species selection

The mitochondrial genomes of six isofemale lines of five species of the buzzatii cluster were assembled for the present study, for which NGS data were available. D. antonietae (Dato) was collected in March 2010 on Martín García Island (Buenos Aires Province, Argentina 34°10′42.12″S 58°15’23.15”W) by J. Hurtado and E. Hasson. D. borborema (Dbrb) was obtained from the Drosophila Species Stock Center (Stock Number: 15081–1281.01, University of California, San Diego, USA) and derived from collections performed in 1974 on Morro do Chapéu (Bahía State, Brazil 11°34′51.98″S 41°07′08.13″W) by M. Wasserman and R.H. Richardson. D. buzzatii (Dbuz) was collected in summer 2010 in Lavalle (Mendoza Province, Argentina 32°37′26.44″S 67°34′15.20″W) by J. Hurtado and E. Hasson. Two D. koepferae (Dkoe) strains; strain B from collections made in Bolivia in December 1982 by A. Fontdevila and A. Ruiz, and strain A collected in Vipos (Tucumán, Argentina 26°28′59″S 65°22′00″W) in February 2010 by J. Hurtado and E. Hasson. D. seriema (Dsei, strain D73C3) derived from collections on Cachoeira do Ferro Doido (Bahía State, Brazil 11°37' 40'' S 41°9' 2'' W) in June 1990 by G. Kuhn and F.M. Sene [56]. The stocks of D. antonietae, D. borborema, D. buzzatii and D. koepferae are available upon request. The rationale of including these D. koepferae strains is motivated by previous protein electrophoresis work showing higher genetic divergence between Bolivian and Argentinian populations than between conspecific populations in other species [45]. In addition, we also included four species of the subgenus Drosophila, for which assembled mitogenomes were available as outgroups in the phylogenetic analyses: D. grimshawi (GenBank: BK006341.1), D. littoralis (GenBank: NC_011596.1), D. virilis (GenBank: BK006340.1) and D. mojavensis (GenBank: BK006339.1).

In silico mtDNA reads extraction

Whole genome sequencing (WGS) and RNA-seq data for D. antonietae, D. borborema and both strains of D. koepferae were generated in our laboratory [34, 50]. For D. seriema and D. buzzatii, mitochondrial reads were retrieved from the Genome sequencing of D. seriema deposited in Sequence Read Archive database (SRA accession ID: ERX2037878) [56] and the D. buzzatii genome project (https://dbuz.uab.cat), respectively. For each species, mitochondrial reads were extracted from genomic and transcriptomic (when available) datasets. Bowtie2 version 2.2.6 [57] was first used with default parameters (end-to-end sensitive mode) to map reads to the mitochondrial genome of D. mojavensis, the closest relative of buzzatii cluster species available, as a reference. Next, only reads that correctly mapped to the reference genome were retained using Samtools version 1.8 [58]. Finally, mapped reads from genomic and transcriptomic datasets were combined to generate a set of only mitochondrial reads.

Mitochondrial reference genome assembly

It is well known that at least 25% of NGS reads are of mitochondrial origin [3]. Therefore, after the mapping process it is possible to attain a coverage ranging from 2000x to more than 20000x for mitogenomes. In order to avoid misassemblies caused by a large number of reads and given the difficulty of determining the coverage (and combination of reads) that recovers the complete mitochondrial genome, we split the reads into several datasets with different coverages by random sampling. Then, a three-step assembly procedure was adopted for these datasets based on recommendations of MITObim package version 1.8 [1]. In the first step, each dataset was employed to build a template by mapping its reads to the mitogenome of D. mojavensis using MIRA assembler [59]. In this way, several templates, based mostly on conserved regions, were built for each species. In the second step, entire mitogenomes were assembled by mapping the complete set of reads to the templates generated in the first step (coverage assembly), individually. This step was performed with the MITObim script and a maximum of ten mapping iterations. Finally, all the different coverage assemblies of the same species were aligned with Clustalw2 version 2.1 [60], and a consensus assembly was then generated considering a sequence representation threshold of 60% (nucleotide mostly represented at each position between assemblies), and not allowing gaps. De novo assemblies for each species, though more fragmented, were aligned to the assemblies obtained as described above and revealed the same gene order along the mitogenomes.

PCR amplification, sanger sequencing and consensus correction

Mitogenome assembly coverage averaged more than 20000x; however, three regions including parts of COI, NADH dehydrogenase subunit 6 (ND6) and large ribosomal RNA (rRNAL) genes showed low read representation in all species, producing miss-assemblies and fragmentation. These regions were PCR-amplified with GO taq Colorless Master Mix by Promega using primers designed for regions conserved across the six mitogenomes assembled in this study (data in S1 Text). PCR amplifications included an initial denaturation at 94°C for 90 s, followed by 25 cycles of denaturation at 94°C for 45 s, annealing at 62°C for 50 s, extension at 72°C for 1 min and a final 4 min extension. PCR fragments were sequenced in both directions on an ABI-3130xl (Genetic Analyzer). Sequences were analyzed and filtered using Mega X software [61] and, finally, merged with the assemblies.

Genome annotation and bioinformatic analyses

The six new assemblies were annotated with the MITOS web server (http://mitos.bioinf.uni-leipzig.de) [62] using the invertebrate mitochondrial genetic code and default parameter settings. The position and orientation of annotations were examined by mapping reads to mitogenomes with Bowtie2 [57] and visualization conducted with IGV ver. 2.4.10 [63]. In addition, nucleotide composition and codon usage were analyzed using MEGA X [61]. A homemade python package (available upon request) was developed to compute the number of pairwise nucleotide differences in the buzzatii cluster, and to visualize its variation along the mitogenomic alignment. Then we used the p-distance as a measured of nucleotide divergence, by dividing the number of nucleotide differences by the total number of nucleotides compared and by the number of pairwise comparisons [61]. Similar p-distance estimates were computed for the D. melanogaster subgroup with the aim to compare divergence patterns along the mitochondrial DNA in the buzzatii cluster with a deeply studied ensemble of species. To this end, one mitogenome of each one of the following species: D. melanogaster (KJ947872.2), D. erecta (BK006335.1), D. simulans (NC_005781), D. sechellia (NC_005780) and D. yakuba (NC_001322.1) were aligned, and nucleotide divergence estimates (p-distance) were obtained as described above. Synonymous (dS) and non-synonymous substitution rates (dN) were also estimated for each mitochondrial protein coding gene (PCG) using PAML 4.8 [64]. These estimates, as well as the ω ratio (dN/dS), were obtained separately for the buzzatii cluster and the melanogaster subgroup sequence alignments. Multiple sequence alignments of each coding gene were obtained with Clustalw2 ver. 2.1 [60].

Phylogenetic analyses

Phylogenetic analyses were conducted considering PCGs, ribosomal genes (rRNAs), transfer RNA genes (tRNAs) and intergenic regions (excluding the control region) of the 6 mitogenomes plus the sequences of the outgroups D. virilis, D. grimshawi, D. littoralis and D. mojavensis (see details in species selection section). An alignment of the ten mitogenomes was performed with Clustalw2 version 2.1 [60]. The flanking sequences that correspond to the control region and portions of the alignment showing abundant gaps were manually removed with Seaview ver. 4 [65]. The final alignment was used as input in PartitionFinder2 [66] to determine the best partition scheme and substitution models, considering separate loci and codon position (in PCGs), which were used in Bayesian Inference and Maximum Likelihood phylogenetic searches. In the Bayesian Inference approach executed with MrBayes ver. 3.2.2 [67], both substitution model and parameter estimates were unlinked. Then, two independent Markov Chain Monte Carlo (MCMC) were run for 30 million generations with three samplings every 1000 generations, for a total of 30,000 trees. Tracer ver. 1.7.1 [68] was used to assess the convergence of the chain mixing, where all parameters had effective sample sizes (ESS) > 200, and 25% of the trees were discarded as burn-in and the remaining trees were used to estimate a consensus tree and the posterior probability of each clade. The consensus tree was plotted and visualized with FigTree ver. 1.4.4 (https://github.com/rambaut/figtree/releases) [69]. Maximum Likelihood searches were performed in 2,000 independent runs using RAxML ver. 8.2.11 [70], applying the rapid hill climb algorithm and the GTR+GAMMA model, considering the partition scheme obtained with PartitionFinder2. Two thousand bootstrap replicates were run to obtain clade frequencies that were plotted onto the tree with highest likelihood. Tree and bootstrap values were visualized with FigTree ver. 1.4.4 [69]. Bayesian Inference searches for each PCG were individually performed to identify correlations with the topology recovered using the complete mitogenome. The GTR-GAMMA model together with the same parameters and evaluation detailed before were applied on each MCMC.

Divergence time estimation

Divergence times were estimated using the same methodology as in Hurtado et al. [50]. Four-fold degenerate third codon sites, i.e. putative neutral sites, of PCGs were extracted from the alignment and Bayesian Inference searches were analyzed with BEAST ver. 1.10.4 [71]. A strict clock was set using a prior for the mutation rate of 6.2x10-07 per year (standard deviation of 1.89x10-07), as was empirically estimated for mitochondrial DNA in Drosophila melanogaster [72]. In addition, a birth-death process with incomplete sampling and a time of 11.3 million years ago -MYA- (CI = ~9.34 to ~13) [26] to the root were defined as tree priors. Two MCMC were produced in 30 million generations with tree sampling every 1000 generations. Tracer [68] was used to evaluate chain convergence, discarding 10% of the all trees (burn-in). The information of the recovered trees was summarized in one tree applying LogCombiner and TreeAnnotator ver. 1.10.4 (available as part of the BEAST package), including the posterior probabilities of the branches, the age of the nodes, and the posterior estimates and HPD limits of the node heights. The target tree was visualized using FigTree [69]. Only D. mojavensis was included as an outgroup in this analysis to minimize problems of among-taxa rate variation given by the large divergence between the buzzatii cluster and the rest of the species already sequenced, together with the lack of time point calibrations and accurate mutation rates.

Results

Mitogenome characterization, nucleotide composition and codon usage

The length of the assembled mitogenomes varied from 14885 to 14899 bp among the six strains reported in this paper. Mitogenomes consisted of a conserved set of 37 genes, including 13 PCGs, 22 tRNAs and 2 rRNAs genes, with order and orientation identical to D. mojavensis. Several short non-coding intergenic regions were also found. Twenty-three genes were found on the heavy strand (+) and fourteen on the light strand (-). Detailed statistics about metrics and composition of the mitogenomes are shown in Table 1.

thumbnail
Table 1. Composition of mitochondrial elements in the species assemblies of the Drosophila buzzatii cluster.

https://doi.org/10.1371/journal.pone.0220676.t001

Overall nucleotide composition in PCGs ranged between 37.6–37.8% A, 37.2–37.9% T, 10.2–10.4% G, and 14.1–14.7% C. The thirteen PCGs were AT-biased as in the entire mitogenome, and the codon usage bias in each gene was greater than 0.50. The most frequently used codons were UUA (Leu), AUU (Ile), and UUU (Phe) in all cases. Codon usage information for each species is shown in Table in S1 Table.

Genetic diversity among mitogenomes

Patterns of divergence (p-distance) along the entire mitogenomes were, overall, very similar for both the buzzatii cluster and the melanogaster subgroup (Fig 2). However, overall divergence was higher in the melanogaster subgroup than in the buzzatii cluster. This difference between both ensembles is probably due to depth of divergence, the melanogaster subgroup comprises pairs of highly diverged species such as D. melanogaster and D. yakuba that split more than 10 MYA [73]. Despite the general similarity in the patterns of divergence, a substantial difference was found in the region encompassing COIII, tRNA-G and ND3 genes. In this region, from position 5000 to 6000, nucleotide diversity was the highest in the melanogaster subgroup, showing an apparent increase represented by two high peaks absent in the buzzatii cluster. Considering genetic divergence within the buzzatii cluster (Fig 3), the lowest value of average pairwise nucleotide divergence was observed for the pair D. borborema and D. seriema (p = 1.91x10-03), while between D. seriema and D. buzzatii divergence was an order of magnitude larger (p = 2.73x10-02). Divergence between D. koepferae strains was surprisingly high (7.14x10-03). The complete set of divergence estimates in the buzzatii cluster is reported in Table in S2 Table. Substitution rates in synonymous (dS) and non-synonymous (dN) sites, and the ω ratios varied among PCGs (Table 2). The dN/dS ratio (ω) varied from 0.003 to 0.060 among PCGs in the buzzatii cluster. The range in the melanogaster subgroup was similar, but with a lower upper bound (0.003–0.018). Two genes, ATP8 and ND2, appear as outliers in the buzzatii cluster (ATP8 and ND2), which apart from these two loci, had lower divergence values than in the melanogaster subgroup. The elevated dN/dS values observed for ATP8 and ND2 in the buzzatii cluster (Table 2) are probably real since the true value of dS may be slightly underestimated due to multiple hits in the melanogaster subgroup (see above), leading to a slight overestimation of the dN/dS ratio. In any case, these results suggest that purifying selection imposes strong constraints in the evolution of mitochondrial genes.

thumbnail
Fig 2. Nucleotide diversity variation along the mitogenome, estimated for a sliding window of 500bp with an overlap of 100bp.

p-distance values for species belonging the buzzatii cluster and the melanogaster subgroup are represented independently.

https://doi.org/10.1371/journal.pone.0220676.g002

thumbnail
Fig 3. Pairwise comparison of nucleotide diversity between species belonging the buzzatii cluster.

The official FlyBase abbreviations for Drosophila species names are shown.

https://doi.org/10.1371/journal.pone.0220676.g003

thumbnail
Table 2. Estimates of non-synonymous (dN) and synonymous (dS) substitutions and their ratio (ω) among species of the buzzatii cluster and the melanogaster subgroup.

https://doi.org/10.1371/journal.pone.0220676.t002

Phylogenetic analyses

The sequences of the 13 PCGs, 22 tRNAs genes, 2 rRNAs genes, and intergenic regions were included in the alignment. Total length of the final matrix encompassing the ten mitogenomes was 15044 characters, from which 1950 were informative sites, 11583 conserved, and 1422 were singletons. Both Maximum Likelihood and Bayesian Inference phylogenetic analyses recovered the same highly supported topology that confirmed the monophyly of the buzzatii cluster (Fig 4). Two main clades can be observed in the tree, one including both D. koepferae strains as sisters to D. buzzatii, and the second, comprising D. antonietae as sister species of the sub-clade formed by D. borborema and D. seriema. The species selected as outgroups were placed as expected, with D. mojavensis as the closest relative of the buzzatii cluster. We also performed a gene tree analysis using all PCGs (S1 Fig). We could only obtain trees for seven genes out of the thirteen PCGs, given the lack of informative sites in the alignments of ATP8, ATP6, ND3, ND4l, COII and COIII. Only two (ND1 and ND5) out of the seven recovered gene trees showed the same topology as the complete mitogenome, while the remaining genes produced three (different) topologies. Trees obtained with CytB and ND4 showed D. buzzatii as sister to the serido sibling set which included D. koepferae. COI and ND2 retrieved trees where D. buzzatii and D. koepferae exchanged positions in the tree, placing D. koepferae as the species closest to the putative ancestor of the cluster. ND6 recovered two clades where D. antonietae was the sister of D. buzzatii and D. koepferae (both strains) in one clade, and the pair D. borborema-D. seriema composed the other.

thumbnail
Fig 4. Phylogenetics hypotheses for the buzzatii cluster based on the entire sequence of the mitogenome (control region not included).

Tree topology recovered by both Maximum Likelihood and Bayesian Inference searches. Bootstrap values and posterior probabilities are indicated at each node.

https://doi.org/10.1371/journal.pone.0220676.g004

Divergence times

PCGs contained 1201 4-fold degenerate sites in the mitogenomes of the buzzatii cluster strains assembled in this study. The tree obtained in the divergence time estimation analysis (Fig 5) was topologically identical to the trees obtained in the phylogenetic analyses using complete mitogenomes (see Fig 4). Divergence time estimations showed that the buzzatii cluster diverged in the Early Pleistocene, 2.11 MYA, and the split with the D. mojavensis common ancestor occurred 10.63 MYA in the Miocene. Our results also indicated that the clade containing D. antonietae, D. borborema and D. seriema, is younger than the clade composed by D. buzzatii and D. koepferae. In addition, the split between D. borborema and D. seriema is quite recent, about ~50,000 years ago, in the Late Pleistocene, even more recent than the split of D. koepferae strains that diverged ~310,000 years ago, in the Middle Pleistocene.

thumbnail
Fig 5. Divergence times for the buzzatii cluster drawn on a Bayesian inference tree.

Numbers on each node are the time estimates. Blue bars represent the 95% confidence intervals of estimates.

https://doi.org/10.1371/journal.pone.0220676.g005

Discussion

The six newly assembled mitochondrial genomes of five cactophilic species of the buzzatii cluster share molecular features with animal mitochondrial genomes sequenced so far [74]. All assembled mitogenomes contain the same set of genes usually found in animal mitochondrial genomes. Gene order and orientation, as well as the distribution of genes on the heavy and light strands were identical to the mitogenome of the closest relative D. mojavensis and other drosophilids [9]. Analysis of overall nucleotide composition of mitogenomes and PCGs revealed the typical AT-bias found in Drosophila mitogenomes. Codon usage is highly biased suggesting that synonymous sites cannot be considered strictly neutral and that some sort of natural selection for translational accuracy governs codon usage [75].

Phylogenetic analyses based either on complete mitogenomes or four-fold degenerate sites (for divergence time estimations), retrieved a well supported tree, suggesting that the cluster is composed by two main clades, one including D. buzzatii and D. koepferae (both strains) and another comprising D. antonietae, D. seriema and D. borborema. Though our present results are consistent with previous work based on single mitochondrial genes [53, 76], they should be considered with caution since we only included a single inbred line as representative of each species, except for D. koepferae. Moreover, phylogenetic relationships depicted by mitogenomes do not agree with phylogenetic studies based on both a small set of nuclear and mitochondrial genes [26] and a large set of nuclear genes -see below- [50]. Interestingly, the topology showing these two clades was only recovered in two (ND1 and ND5) out of seven trees based on individual PCGs, and the remaining gene trees produced either a novel topology or a topology consistent with the phylogeny reported in Hurtado et al. [50].

The lack of recombination causes mitochondrial DNA to be inherited as a unit, so trees recovered with individual mitochondrial genes are expected to share the same topology and to be consistent with the trees obtained with complete mitogenomes. On the other hand, our results suggest that individual genes not only produce different topologies but also a poor resolution of phylogenetic relationships. Such inconsistencies between complete mitogenomes and gene trees in phylogenetic estimation may result from inaccurate reconstruction or from real differences among gene trees. The first possible explanation is the simple fact that numbers of informative sites within a single locus are insufficient to accurately estimate phylogenetic relationships, particularly in groups of recently diverged species: overall support for gene trees was poorer than for the tree based on complete mitogenomes. Second, heterogeneity in evolutionary rates among genes and/or differences in selective constraints along the mitogenome can also account for these inconsistencies [17, 7779] consistent with our observations of substantial variation in synonymous and nonsynonymous rates as well as ω ratios across PCGs. In addition, variation among oxidative phosphorylation complexes in the buzzatii cluster was high. The ND complex was, on average, less constrained than the ATP complex, cytochrome b (CytB) and cytochrome oxidase complex (COI, COII & COIII), consistent with results reported in the melanogaster subgroup [9, 80]. Another factor that may lead to biased tree construction, particularly relevant for mitochondrial genes characterized by high substitution rates, is substitutional saturation [81]. A priori, saturation should not be problematic in recently diverged species, like the buzzatii cluster, however, saturation may be problematic in the estimation of divergence relative to the outgroup and, thus, for phylogenetic inference. The closest outgroup to the buzzatii cluster employed in our study was the mulleri complex species D. mojavensis. Available evidence suggest that these complexes diverged ~10 MYA [26], but possibly more recently, ca 5.5 MYA [50] suggesting that substitution saturation may lead to inaccurate phylogenetic reconstruction.

In this context, a recent report investigating the effect of using individual genes, subsets of genes, complete mitogenomes and different partitioning schemes on tree topology suggested a framework to interpret the results of mitogenomic phylogenetic studies [11]. The authors concluded that trees obtained with complete mitogenomes reach the highest phylogenetic performance and reliability than single genes or subsets of genes. Therefore, we consider that phylogenetic relationships inferred from complete mitogenomes reflect the evolutionary history of, at least, mitogenomes.

The phylogenetic relationships depicted by our mitogenomic approach are incongruent with a recent study based on transcriptomic data [50]. Based on a concatenated matrix of 813 kb uncovering 761 gene regions, the authors obtained a well-supported topology in which D. koepferae appears phylogenetically closer to D. antonietae and D. borborema than to D. buzzatii, placing D. buzzatii alone as sister to the rest of the cluster. This topology agrees with male genital morphology, cytological and molecular phylogenetic evidence [26, 35, 55]. Nevertheless, the pattern of cladogenesis of the trio D. koepferae-D. borborema-D. antonietae could not be fully elucidated since a nuclear gene tree analysis yielded ambiguous results. As a matter of fact, the analysis of the 761 gene trees reported showed that about one third of the genes supported each one of the three possible topologies for the trio D. koepferae-D. antonietae-D. borborema indicating a hard polytomy [50]. In contrast, the early separation of D. buzzatii from the serido sibling set is supported by 97% of the gene trees and, surprisingly, none of the gene trees recovered the clade including D. buzzatii and D. koepferae as the sister group of the clade involving D. antonietae and D. borborema [50] as suggested by the present mitogenomic approach. Such mitonuclear discordance has been reported in several animal species. A recent review lists several examples in animals [82]. Likewise, the literature in this respect is abundant in the genus Drosophila. Well-known cases are D. pseudoobscura and D. persimilis [83]; D. santomea and D. yakuba [84]; and D. simulans and D. mauritiana [85]. Mitonuclear discordance may be caused by incomplete lineage sorting (ILS) and/or introgressive hybridization. These two factors do not equally affect mitochondrial and nuclear genomes because ILS is more likely for nuclear genes, especially when the ancestral effective population size of recently diverged species was large [86, 87]. Introgressive hybridization is expected to be prevalent in mitochondrial genomes given its lower effective population size [88]. If we accept that the topology based on nuclear genes is representative of the species-history (see also [49]), the greater similarity between D. buzzatii and D. koepferae mitogenomes is suggestive of gene flow between these largely sympatric species [35]. Thus, we suggest that D. buzzatii and D. koepferae lineages initially separated but then exchanged genes via fertile F1 females (males were likely sterile as expected due to Haldane’s Rule) before finally separating less than 1.5 MYA. Not only the more recent mitogenomic ancestry is suggestive of gene exchange, also traces of introgressive hybridization can still be detected in nuclear genomes [50].

In fact, phylogenetic, population genetic and experimental hybridization studies suggest a significant role of introgression in the evolutionary history of the buzzatii cluster. Phylogeographic studies revealed discordances between mitochondrial markers and genital morphology in areas of sympatry between species [53]. Likewise, interspecific gene flow has been invoked to account for shared nucleotide polymorphisms in nuclear genes in D. buzzatii and D. koepferae that cannot be accounted by ILS [89, 90]. Moreover, experimental hybridization studies have shown that several species of the buzzatii cluster can be successfully crossed, producing fertile hybrid females that can be backcrossed to both parental species. Interestingly, D. koepferae can be successfully crossed with D. antonietae, D. borborema, D. buzzatii and D. serido [45, 48, 9195].

Our estimates of divergence times are in conflict with most previous studies. In general, previous estimates, based on individual genes or a few genes (either mitochondrial or nuclear) suggested an older origin of the cluster and deeper splitting times within the cluster when compared to the estimates based on transcriptomes and mitogenomes. In effect, Gómez & Hasson [89] and Oliveira et al. [26] dated the split of D. buzzatii from the remaining species of the cluster at 4.6 MYA, respectively, whereas Manfrin et al.’s estimates [48] were even older, from 3 to 12 MYA for the most recent to the more ancient split. In contrast, putting apart the divergence time of the clade D. buzzatii-D. koepferae for the reasons discussed above, the radiation of the remaining three species seems to be extremely recent, less than 1 MYA using mitochondrial genomes (Fig 5), which are similar to estimates based on transcriptomes [50]. However, it is worth mentioning that divergence times estimated in the present paper and by Hurtado et al. [50] may be biased downwards since both are based on empirical mutation rates for nuclear and mitochondrial genes, respectively, calculated for over 200 generations in D. melanogaster [72]. Thus, these results should be interpreted with caution in the light of evidence suggesting not only the time-dependence of molecular evolutionary rates, but also that mutation rates obtained using pedigrees and laboratory mutation-accumulation lines often exceed long-term substitution rates by an order of magnitude or more [79].

Even though divergence times estimates obtained in this study cannot be entirely compared to assessments based on nuclear genomic data and individual nuclear genes, given the uncertainty of tree topology, they concur in that species of the buzzatii cluster apparently emerged during the Late Pleistocene in association with Quaternary climate fluctuations [49, 50, 76]. Moreover, in view of the obligate ecological association between buzzatii cluster species and cacti, the so-called Pleistocene “refuge hypothesis” is a suitable explanation for the diversification in this group in active cladogenesis. This hypothesis suggests that Pleistocene glacial cycles successively generated isolated patches of similar habitats across which populations may have diverged into species [96, 97].

Available paleo-climatic evidence, consistent with the Pleistocene “refuge hypothesis”, can also account for the relatively deep intraspecific divergence between Bolivian and Argentinian D. koepferae strains. Because Quaternary topographical patterns in the Central Andes have remained unchanged for the last 2–3 MYA, a plausible explanation for this late Pleistocene vicariant event is related with glacial-interglacial cycles [98]. Although the validity of the Pleistocene “refuge hypothesis” is controversial (cf. [99]) and few studies addressed specific hypotheses on how the Quaternary glacial-interglacial cycles impacted species diversification [100], our divergence time estimates between Bolivian and Argentinian D. koepferae suggest a role of climatic oscillations as a factor of ecogeographical isolation in the Central Andes during the Pleistocene. Moreover, paleo-climatological evidence suggest that the area inhabited by D. koepferae has been exposed to substantial climatic variations on timescales of 103–105 years related with glacial-interglacial cycles. Thus, Andean north-south exchanges may have been alternately favored or disfavored by these Quaternary climatic oscillations. In fact, the estimated age of the vicariant event between the D. koepferae strains is tantalizingly coincident with the coldest phase of the Marine Isotopic Stage (MIS) 10, which corresponds to a glacial period that ended about 337,000 years ago [101]. The coldest period of the MIS 10, recorded in global air and sea surface temperature and also the lowest atmospheric CO2 levels, occurred ca 355,000 years, well within the confidence interval of our divergence time estimated between D. koepferae strains. On a global scale, glacial periods are primarily reflected in a lowering of air temperature but also in altered patterns of precipitation in the both sides of the Central Andes [102] which were in turn the main drivers of vegetation changes [103] including the appearance of South American columnar cacti [104]. Besides the impact on air temperature, periods of ice advance in the Central Andes generally were periods of negative water balance in the Pacific coastal regions west to the Central Andes [105], and a positive water balance in the Central Andes, as evidenced by deeper and fresher conditions in Lake Titicaca [106] (see S2 Fig). Thus, during the colder and wetter phases of the MIS 10 in the Central Andes, species distributions may have suffered a general contraction towards the southern and northern lowland warmer refugia between 1000–2000 m, whereas a general worsening condition occurred in higher western elevations. Northern and southern refugia were probably separated by a gap of low suitability represented by the steep gradient of the eastern flank of Eastern Andes between 22–24°S, which represents today a region of strong W-E precipitation gradient. The MIS 10 glacial cycle has a particular structure since it does not have a pronounced interstadial (relative warmer) conditions in the mid-cycle [107], providing a prolonged, effective “soft” dispersal barrier that affected the distribution of D. koepferae.

Finally, future analyses including the mitogenomes of the other Brazilian species D. gouveai and D. serido and several mitogenomes of each species are needed to achieve a more complete understanding of the evolutionary history of the cluster. Such comparative analysis including the complete mitogenomes of all buzzatii cluster species will help to disentangle the intricate relationships in this group.

Supporting information

S1 Text. Set of primers used to sequence each gene.

https://doi.org/10.1371/journal.pone.0220676.s001

(DOCX)

S1 Table. Codon usage for each mitogenome of the buzzatii cluster species.

https://doi.org/10.1371/journal.pone.0220676.s002

(XLSX)

S2 Table. Genetic divergence among species of the buzzatii cluster.

Estimates are shown for each pairwise comparison between species.

https://doi.org/10.1371/journal.pone.0220676.s003

(XLSX)

S1 Fig. Phylogenetic hypotheses for the buzzatii cluster species recovered by each mitochondrial gene using Bayesian Inference searches.

https://doi.org/10.1371/journal.pone.0220676.s004

(TIF)

S2 Fig. Paleoclimatic records of the last 500,000 years.

Ages in the top are indicated as 103 years (kyrs). Gradated shading area indicates divergence age estimates. Marine Isotope Stages (MIS) are labeled according to Lisiecki and Raymo [101]. Shaded vertical areas correspond to glacial periods whereas white areas correspond to interglacials or interstadials. Glacial periods correspond to cold and dry conditions in the western slopes of the Western Andes, and cold and wetter conditions in the eastern slopes of the Eastern Andes and the Altiplano. A. Globally-averaged surface air temperature anomaly reconstructed from proxy and model data for the last eight glacial cycles [108]. B. CO2 concentration based on Vostok Ice Core data [109]. C. Iron accumulation rates (AR Fe) reflecting changes in terrigenous sediment input to ODP Site 1239D, Equatorial Pacific [110]. D. % of CaCO3from Site LT01-2B indicating changes in water balance at Lake Titicaca Basin, Bolivia (modified from [106]).

https://doi.org/10.1371/journal.pone.0220676.s005

(TIF)

Acknowledgments

The authors wish to thank D. Rand, an anonymous reviewer and the academic editor W.J. Etges for useful comments and constructive criticisms that helped to improve previous versions of the manuscripts. NNM is a CONICET PhD student. JM, JH, FA and EH are members of Carrera del Investigador Científico y Tecnológico (CIC) of CONICET.

References

  1. 1. Hahn C, Bachmann L, Chevreux B. Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—A baiting and iterative mapping approach. Nucleic Acids Res. 2013;41(13): e129–e129. pmid:23661685
  2. 2. Tian Y, Smith DR. Recovering complete mitochondrial genome sequences from RNA-Seq: A case study of Polytomella non-photosynthetic green algae. Mol Phylogenet Evol. 2016;98: 57–62. pmid:26860338
  3. 3. Smith DR. The past, present and future of mitochondrial genomics: Have we sequenced enough mtDNAs? Brief Funct Genomics. 2016;15(1): 47–54. pmid:26117139
  4. 4. Osigus HJ, Eitel M, Bernt M, Donath A, Schierwater B. Mitogenomics at the base of Metazoa. Mol Phylogenet Evol. 2013;69: 339–351. pmid:23891951
  5. 5. Roos J, Aggarwal RK, Janke A. Extended mitogenomic phylogenetic analyses yield new insight into crocodylian evolution and their survival of the Cretaceous-Tertiary boundary. Mol Phylogenet Evol. 2007;45: 663–673. pmid:17719245
  6. 6. Finstermeier K, Zinner D, Brameier M, Meyer M, Kreuz E, Hofreiter M, et al. A Mitogenomic Phylogeny of Living Primates. PLoS ONE. 2013;8(7): e69504. pmid:23874967
  7. 7. Saitoh K, Sado T, Mayden RL, Hanzawa N, Nakamura K, Nishida M, et al. Mitogenomic Evolution and Interrelationships of the Cypriniformes (Actinopterygii: Ostariophysi): The First Evidence Toward Resolution of Higher-Level Relationships of the World’s Largest Freshwater Fish Clade Based on 59 Whole Mitogenome Sequences. J Mol Evol. 2006;63: 826–841. pmid:17086453
  8. 8. Zhang P, Liang D, Mao RL, Hillis DM, Wake DB, Cannatella DC. Efficient Sequencing of Anuran mtDNAs and a Mitogenomic Exploration of the Phylogeny and Evolution of Frogs. Mol Biol Evol. 2013;30(8): 1899–1915. pmid:23666244
  9. 9. Montooth KL, Abt DN, Hofmann JW, Rand DM. Comparative genomics of Drosophila mtDNA: novel features of conservation and change across functional domains and lineages. J Mol Evol. 2009;69(1): 94. pmid:19533212
  10. 10. Cameron SL. Insect Mitochondrial Genomics: Implications for Evolution and Phylogeny. Annu Rev Entomol. 2001;59(1): 95–117.
  11. 11. Duchêne S, Archer FI, Vilstrup J, Caballero S, Morin PA. Mitogenome Phylogenetics: The Impact of Using Single Regions and Partitioning Schemes on Topology, Substitution Rate and Divergence Time Estimation. PLoS ONE. 2011;6(11): e27138. pmid:22073275
  12. 12. Rohland N, Malaspinas AS, Pollack JL, Slatkin M, Matheus P, Hofreiter M. Proboscidean mitogenomics: chronology and mastodon as outgroup. PLoS Biol. 2007;5: e207. pmid:17676977
  13. 13. Willerslev E, Gilbert T, Binladen J, Ho SYW, Campos PF, Ratan A, et al. Analysis of complete mitochondrial genomes from extinct and extant Rhinoceroses reveals lack of phylogenetic resolution. BMC Evol Biol. 2009;9: 95. pmid:19432984
  14. 14. Knaus BJ, Cronn R, Liston A, Pilgrim K, Schwartz MK. Mitochondrial genome sequences illuminate maternal lineages of conservation concern in a rare carnivore. BMC Ecol. 2011;11: 10. pmid:21507265
  15. 15. Luo A, Zhang A, Ho SYW, Xu W, Zhang Y, Shi W, et al. Potential efficacy of mitochondrial genes for animal DNA barcoding: a case study using eutherian mammals. BMC Genomics. 2011;12: 84. pmid:21276253
  16. 16. Pacheco MA, Battistuzzi FU, Lentino M, Aguilar R, Kumar S, Escalante AA. Evolution of modern birds revealed by mitogenomics: timing the radiation and origin of major orders. Mol Biol Evol. 2011;28: 1927–1942. pmid:21242529
  17. 17. Balloux F. The worm in the fruit of the mitochondrial DNA tree. Heredity. 2010;104: 419–420. pmid:19756036
  18. 18. Pamilo P, Nei M. Relationships between gene trees and species trees. Mol Biol Evol. 1988;5(5): 568–583. pmid:3193878
  19. 19. Maddison WP. Gene trees in species trees. Syst Biol. 1997;46(3): 523–536.
  20. 20. Zachos FE. Gene trees and species trees–mutual influences and interdependences of population genetics and systematics. J Zool Syst Evol Res. 2009;47(3): 209–218.
  21. 21. Bächli G. [Internet]. TaxoDros: the database on taxonomy of Drosophilidae, v. 1.04. Database 2011/1. Available from: https://www.taxodros.uzh.ch/.
  22. 22. Markow TA, O’Grady P. Reproductive ecology of Drosophila. Funct Ecol. 2008;22(5): 747–759.
  23. 23. Markow TA, O’Grady P. Drosophila: A guide to species identification and use. Elsevier; 2005.
  24. 24. O’Grady PM, DeSalle R. Phylogeny of the genus Drosophila. Genetics. 2018;209(1): 1–25. pmid:29716983
  25. 25. Markow TA. Host use and host shifts in Drosophila. Curr Opin Insect Sci. 2019;31: 139–145. pmid:31109667
  26. 26. Oliveira DCSG, Almeida FC, O’Grady PM, Armella MA, DeSalle R, Etges WJ. Monophyly, divergence times, and evolution of host plant use inferred from a revised phylogeny of the Drosophila repleta species group. Mol Phylogenet Evol. 2012;64(3): 533–544. pmid:22634936
  27. 27. Morales-Hojas R, Vieira J. Phylogenetic Patterns of Geographical and Ecological Diversification in the Subgenus Drosophila. PLoS ONE. 2012;7(11): e49552. pmid:23152919
  28. 28. Heed WB. Ecology and genetics of Sonoran desert Drosophila. In: Brussard PF, editors. Ecological Genetics: The Interface. Proceedings in Life Sciences. Springer, New York, NY; 1978. pp. 109–126.
  29. 29. Barker JS, Starmer W. Ecological Genetics and Evolution: The Cactus-Yeast-Drosophila Model System. Academic Pr; 1982.
  30. 30. Heed WB, Mangan RL. Community ecology of Sonoran Desert Drosophila. In: Asburner M., Carson H., Thompson J. N., editors. The genetics and biology of Drosophila. Academic, London; 1986. pp. 311–345.
  31. 31. Hasson E, Naveira H, Fontdevila A. The Breeding Sites of Argentinean Cactophilic Species of the Drosophila-Mulleri Complex (Subgenus Drosophila-Repleta Group). Rev. Chil. Hist. Nat. 1992;65(3): 319–326.
  32. 32. Fogleman JC, Danielson PB. Chemical interactions in the Cactus-Microorganism-Drosophila Model System of the Sonoran Desert. Am Zool, 2001;41(4): 877–889.
  33. 33. Guillén Y, Rius N, Delprat A, Williford A, Muyas F, Puig M, et al. Genomics of ecological adaptation in cactophilic Drosophila. Genome Biol Evol. 2014;7(1): 349–366. pmid:25552534
  34. 34. De Panis DN, Padró J, Furió Tarí P, Tarazona S, Milla Carmona PS, Soto IM, et al. Transcriptome modulation during host shift is driven by secondary metabolites in desert Drosophila. Mol Ecol. 2016;25(18): 4534–4550. pmid:27483442
  35. 35. Hasson E, De Panis D, Hurtado J, Mensch J. Host plant adaptation in cactophilic species of the Drosophila buzzatii cluster: fitness and transcriptomics. J Hered. 2019;110(1): 46–57. pmid:30107510
  36. 36. Throckmorton LH. The Phylogeny, Ecology, and Geography of Drosophila. In: King RC, editors. Plenum Publishing Corporation, New York, New York; 1975. vol. 3, pp. 421–469.
  37. 37. Wasserman M. Evolution in the repleta group. In: Ashburner M, Carson HL, Thompson JN, editors. The genetics and Biology of Drosophila. Academic Press, London; 1982. pp. 61–139.
  38. 38. Vilela CA. A revision of the Drosophila species group. (Diptera-Drosophilidae). Rev Bras Entomol. 1983;27, 1±114.
  39. 39. Markow TA, O’Grady P. Drosophila: A guide to species identification and use. Elsevier; 2006.
  40. 40. Drosophila 12 Genomes Consortium. Evolution of genes and genomes on the Drosophila phylogeny. Nature. 2007;450(7167): 203–218. pmid:17994087
  41. 41. Ruiz A, Wasserman M. Evolutionary cytogenetics of the drosophila buzzatii species complex. Heredity (Edinb). 1993;70(6): 582–596.
  42. 42. Tidon-Sklorz R, Sene FM. Two new species of the Drosophila serido sibling set (Diptera, Drosophilidae). Iheringia Ser. Zool. Iheringia. 2001;(90): 141–146.
  43. 43. Vilela CR, Sene FM. Two new Neotropical species of the repleta group of the genus Drosophila (Diptera, Drosophilidae). Pap Avulsos Zool. 1977;30(20): 295–299.
  44. 44. Patterson JT, Wheeler MR. Description of new species of the subgenera Hirtodrosophila and Drosophila. University of Texas. 1942.
  45. 45. Fontdevila A, Pla C, Hasson E, Wasserman M, Sanchez A, Naveira H, et al. Drosophila koepferae: a new member of the Drosophila serido (Diptera: Drosophilidae) superspecies taxon. Ann Entomol Soc Am. 1988;81(3): 380–385.
  46. 46. Tidon-Sklorz R, De Melo Sene F. Drosophila seriema n. sp.: new member of the Drosophila serido (Diptera: Drosophilidae) superspecies taxon. Ann Entomol Soc Am. 1995;88(2): 139–142.
  47. 47. Fontdevila A. Founder Effects in Colonizing Populations: The Case of Drosophila buzzatii. In: Fontdevila A, editors. Evolutionary Biology of Transient Unstable Populations. Springer, New York, NY; 1989. pp. 74–95.
  48. 48. Manfrin MH, Sene FM. Cactophilic Drosophila in South America: A model for evolutionary studies. Genetica, 2006;126(1–2): 57–75. pmid:16502085
  49. 49. Barrios-Leal DY, Neves-Da-Rocha J, Manfrin MH. Genetics and Distribution Modeling: The Demographic History of the Cactophilic Drosophila buzzatii Species Cluster in Open Areas of South America. J Hered. 2019;110(1): 22–33. pmid:30252085
  50. 50. Hurtado J, Almeida F, Revale S, Hasson E. Revised phylogenetic relationships within the Drosophila buzzatii species cluster (Diptera: Drosophilidae: Drosophila repleta group) using genomic data. Arthropod Systematics and Phylogeny. 2019;77(2): 239–250.
  51. 51. Hasson E, Soto IM, Carreira VP, Corio C, Soto EM, Betti M. Host plants, fitness and developmental instability in a guild of cactophilic species of the genus Drosophila. In: Santos EB, editors. Ecotoxicology research developments. Nova Science Publishers, Inc; 2009. pp. 89–109.
  52. 52. Ruiz A, Cansian AM, Kuhn GC, Alves MA, Sene FM. The Drosophila serido speciation puzzle: putting new pieces together. Genetica. 2000;108(3): 217–227. pmid:11294608
  53. 53. Manfrin MH, de Brito ROA, Sene FM. Systematics and Evolution of the Drosophila buzzatii (Diptera: Drosophilidae) Cluster Using mtDNA. Ann Entomol Soc Am. 2001;94(3): 333–346.
  54. 54. Franco FF, Silva-Bernardi ECC, Sene FM, Hasson ER, Manfrin MH. Intra‐and interspecific divergence in the nuclear sequences of the clock gene period in species of the Drosophila buzzatii cluster. J Zool Syst Evol Res. 2010;48(4): 322–331.
  55. 55. Rodríguez-Trelles F, Alarcón L, Fontdevila A. Molecular evolution and phylogeny of the buzzatii complex (Drosophila repleta group): A maximum-likelihood approach. Mol Biol Evol. 2000;17(7): 1112–1122. pmid:10889224
  56. 56. de Lima LG, Svartman M, Kuhn GCS. Dissecting the Satellite DNA Landscape in Three Cactophilic Drosophila Sequenced Genomes. G3 (Bethesda). 2017;7(8): 2831–2843.
  57. 57. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9(4): 357–359. pmid:22388286
  58. 58. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The sequence alignment/map format and SAMtools. Bioinformatics. 2009;25(16): 2078–2079. pmid:19505943
  59. 59. Chevreux B, Pfisterer T, Drescher B, Driesel AJ, Müller WEG, Wetter T. et al. Using the miraEST assembler for reliable and automated mRNA transcript assembly and SNP detection in sequenced ESTs. Genome Res. 2004;14(6): 1147–1159. pmid:15140833
  60. 60. Sievers F, Higgins DG. Clustal omega. Curr Protoc Bioinformatics. 2014;48(1): 3–13.
  61. 61. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6): 1547–1549. pmid:29722887
  62. 62. Bernt M, Donath A, Jühling F, Externbrink F, Florentz C, Fritzsch G, et al. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol Phylogenet Evol. 2013;69(2): 313–319. pmid:22982435
  63. 63. Robinson JT, Thorvaldsdóttir H, Winckler W, Guttman M, Lander ES, Getz G, et al. Integrated genomics viewer. Nat Biotechnol. 2011;29(1): 24–26. pmid:21221095
  64. 64. Yang Z. PAML 4: a program package for phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24: 1586–1591. pmid:17483113
  65. 65. Gouy M, Guindon S, Gascuel O. Sea view version 4: A multiplatform graphical user interface for sequence alignment and phylogenetic tree building. Mol Biol Evol. 2010;27(2): 221–224. pmid:19854763
  66. 66. Lanfear R, Frandsen PB, Wright AM, Senfeld T, Calcott B. Partitionfinder 2: New methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol Biol Evol. 2017;34(3): 772–773. pmid:28013191
  67. 67. Ronquist F, Huelsenbeck JP. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19(12): 1572–1574. pmid:12912839
  68. 68. Rambaut A, Drummond AJ, Xie D, Baele G, Suchard MA. Posterior Summarization in Bayesian Phylogenetics Using Tracer 1.7. Syst Biol. 2018;67(5): 901–904. pmid:29718447
  69. 69. Rambaut A. FigTree: Tree Figure Drawing Tool [software]. 2007. Available online from: http://tree.bio.ed.ac.uk/software/figtree.
  70. 70. Stamatakis A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9): 1312–1313. pmid:24451623
  71. 71. Suchard MA, Lemey P, Baele G, Ayres DL, Drummond AJ, Rambaut A. Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10. Virus Evol. 2018;4: vey016. pmid:29942656
  72. 72. Haag-Liautard C, Coffey N, Houle D, Lynch M, Charlesworth B, Keightley PD. Direct estimation of the mitochondrial DNA mutation rate in Drosophila melanogaster. PLoS Biol. 2008;6(8): 1706–1714.
  73. 73. Tamura K, Subramanian S, Kumar S. Temporal Patterns of Fruit Fly (Drosophila) Evolution Revealed by Mutation Clocks. Mol Biol Evol. 2014;21: 36–44.
  74. 74. D'Onorio de Meo P, D'Antonio M, Griggio F, Lupi R, Borsani M, Pavesi G, et al. MitoZoa 2.0: a database resource and search tools for comparative and evolutionary analyses of mitochondrial genomes in Metazoa. Nucleic Acids Res. 2011;40(D1): D1168–D1172.
  75. 75. Stoletzki N, Eyre-Walker A. Synonymous codon usage in Escherichia coli: selection for translational accuracy. Mol Biol Evol. 2007;24: 374–81. pmid:17101719
  76. 76. Franco FF, Manfrin MH. Recent demographic history of cactophilic Drosophila species can be related to Quaternary palaeoclimatic changes in South America. J Biogeogr. 2013;40(1): 142–154.
  77. 77. Subramanian S. Temporal trails of natural selection in human mitogenomes. Mol Biol Evol. 2009;26: 715–717. pmid:19150805
  78. 78. Subramanian S, Denver DR, Millar CD, Heupink T, Aschrafi A, Emslie SD, et al. High mitogenomic evolutionary rates and time dependency. Trends Genet. 2009;25: 482–486. pmid:19836098
  79. 79. Ho SY, Lanfear R, Bormham L, Phillips MJ, Soubrier J, Rodrigo AG, et al. Time-dependent rates of molecular evolution. Mol Ecol. 2011;20: 3087–3101. pmid:21740474
  80. 80. Ballard JWO. Comparative genomics of mitochondrial DNA in Drosophila simulans. J Mol Evol. 2000;51(1): 64–75. pmid:10903373
  81. 81. Brown WM, George M, Wilson AC. Rapid evolution of animal mitochondrial DNA. Proc Natl Acad Sci U S A. 1979;76(4): 1967–1971. pmid:109836
  82. 82. Toews DP, Brelsford A. The biogeography of mitochondrial and nuclear discordance in animals. Molecular Ecology. 2012;21(16): 3907–3930. pmid:22738314
  83. 83. Powell JR. Interspecific cytoplasmic gene flow in the absence of nuclear gene flow: evidence from Drosophila. Proc Natl Acad Sci U S A. 1983;80(2): 492–495. pmid:6300849
  84. 84. Bachtrog D, Thornton K, Clark A, Andolfatto P. Extensive introgression of mitochondrial DNA relative to nuclear genes in the Drosophila yakuba species group. Evolution (N Y). 2006;60(2): 292–302.
  85. 85. Aubert J, Solignac M. Experimental evidence for mitochondrial DNA introgression between Drosophila species. Evolution (NY). 1990;44(5): 1272–1282.
  86. 86. Wong A, Jensen JD, Pool JE, Aquadro CF. Phylogenetic incongruence in the Drosophila melanogaster species group. Mol Phylogenet Evol. 2007;43(3): 1138–1150. pmid:17071113
  87. 87. Chan KMA, Levin SA. Leaky prezygotic isolation and porous genomes: rapid introgression of maternally inherited DNA. Evolution (NY). 2005;59: 720–729.
  88. 88. Keck BP, Near TJ. Geographic and temporal aspects of mitochondrial replacement in Nothonotus darters (Teleostei: Percidae: Etheostomatinae). Evolution. 2010;64(5): 1410–428. pmid:19930456
  89. 89. Gómez GA, Hasson E. Transpecific polymorphisms in an inversion linked esterase locus in Drosophila buzzatii. Mol Biol Evol. 2003;20(3): 410–423. pmid:12644562
  90. 90. Piccinali R, Aguadé M, Hasson E. Comparative molecular population genetics of the Xdh locus in the cactophilic sibling species Drosophila buzzatii and D. koepferae. Mol Biol Evol. 2004;21(1): 141–152. pmid:14595098
  91. 91. Madi-Ravazzi L, Bicudo HE, Manzato JA. Reproductive compatibility and chromosome pairing in the Drosophila buzzatii complex. Cytobios. 1997;89(356): 21–30. pmid:9297813
  92. 92. Machado LPB, Madi-Ravazzi L, Tadei WJ. Reproductive relationships and degree of synapsis in the polytene chromosomes of the Drosophila buzzatii species cluster. Braz J Biol. 2006;66(1B): 279–293. pmid:16710520
  93. 93. Soto IM, Carreira VP, Fanara JJ, Hasson E. Evolution of male genitalia: environmental and genetic factors affect genital morphology in two Drosophila sibling species and their hybrids. BMC Evol Biol. 2007;7(1): 77.
  94. 94. Soto EM, Soto IM, Carreira VP, Fanara JJ, Hasson E. Host‐related life history traits in interspecific hybrids of cactophilic Drosophila. Entomol Exp Appl. 2008;126(1): 18–27.
  95. 95. Iglesias PP, Hasson E. The role of courtship song in female mate choice in South American Cactophilic Drosophila. PLoS ONE. 2017;12(5): e0176119. pmid:28467464
  96. 96. Haffer J. Speciation in Amazonian forest birds. Science. 1969;165: 131–137. pmid:17834730
  97. 97. Endler JA. Problems in distinguishing historical from ecological factors in biogeography. Am Zool. 1982;22(2): 441–452.
  98. 98. Rull V. Neotropical biodiversity: timing and potential drivers. Trends Ecol Evol. 2011;26(10): 508–513. pmid:21703715
  99. 99. Hoorn C, Wesselingh FP, Ter Steege H, Bermudez MA, Mora A, Sevink J, et al. Response to Origins of Biodiversity. Science 2011;331: 399–400.
  100. 100. Lagomarsino LP, Condamine FL, Antonelli A, Mulch A, Davis CC. The abiotic and biotic drivers of rapid diversification in Andean bellflowers (Campanulaceae). New Phytol. 2016;210: 1430–1442. pmid:26990796
  101. 101. Lisiecki LE, Raymo ME. A Pliocene-Pleistocene stack of globally distributed benthic stable oxygen isotope records. Paleoceanography. 2005;20: 1–17.
  102. 102. Mosblech NA, Bush MB, Gosling WD, Hodell D, Thomas L, Van Calsteren P. North Atlantic forcing of Amazonian precipitation during the last ice age. Nat Geosci. 2012;5(11): 817.
  103. 103. Gosling WD, Bush MB, Hanselman JA, Chepstow-Lusty A. Glacial-interglacial changes in moisture balance and the impact on vegetation in the southern hemisphere tropical Andes (Bolivia/ Peru). Palaeogeogr Palaeoclimatol Palaeoecol. 2008;259: 35–50.
  104. 104. Quipildor VB, Kitzberger T, Ortega-Baes P, Quiroga MP, Premoli AC. Regional climate oscillations and local topography shape genetic polymorphisms and distribution of the giant columnar cactus Echinopsis terscheckii in drylands of the tropical Andes. J Biogeogr. 2017;45: 116–126.
  105. 105. Zhang S, Li T, Chang F, Yu Z, Xiong Z, Wang H. Correspondence between the ENSO-like state and glacial-interglacial condition during the past 360 kyr. Chin. J. Oceanol. Limnol. 2016;35(5): 1018–1031.
  106. 106. Fritz SC, Baker PA, Tapia P, Spanbauer T, Westover K. Evolution of the Lake Titicaca basin and its diatom flora over the last ~370,000 years. Palaeogeogr Palaeoclimatol Palaeoecol. 2012;317–318: 93–103.
  107. 107. Hughes PD, Gibbard PL. Global glacier dynamics during 100 ka Pleistocene glacial cycles. Quat Res. 2018;90(1): 222–243.
  108. 108. Friedrich T, Timmermann A, Tigchelaar M, Timm OE, Ganopolski A. Nonlinear climate sensitivity and its implications for future greenhouse warming. Sci Adv. 2016;2(11): e1501923. pmid:28861462
  109. 109. Petit JR, Jouzel J, Raynaud D, Barkov NI, Barnola JM, Basile I, et al. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature. 1999;399: 429–436.
  110. 110. Rincón-Martínez D, Lamy F, Contreras S, Leduc G, Bard E, Saukel C, et al. More humid interglacials in Ecuador during the past 500 kyr linked to latitudinal shifts of the equatorial front and the Intertropical Convergence Zone in the eastern tropical Pacific. Paleoceanography. 2010;25: PA2210.