Skip to main content
Advertisement
  • Loading metrics

An investigation of Y chromosome incorporations in 400 species of Drosophila and related genera

  • Eduardo G. Dupim,

    Roles Data curation, Investigation, Writing – original draft, Writing – review & editing

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

  • Gabriel Goldstein,

    Roles Investigation, Writing – review & editing

    Affiliations Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil, Departamento de Genética e Biologia Evolutiva, Universidade de São Paulo, São Paulo, SP, Brazil

  • Thyago Vanderlinde,

    Roles Data curation, Investigation, Writing – review & editing

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

  • Suzana C. Vaz,

    Roles Data curation, Resources, Writing – review & editing

    Affiliations Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil, Departamento de Genética e Biologia Evolutiva, Universidade de São Paulo, São Paulo, SP, Brazil

  • Flávia Krsticevic,

    Roles Investigation, Writing – review & editing

    Affiliations Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil, CIFASIS, CONICET, Rosario, Santa Fe, Argentina

  • Aline Bastos,

    Roles Investigation

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

  • Thadeo Pinhão,

    Roles Investigation

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

  • Marcos Torres,

    Roles Investigation

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

  • Jean R. David,

    Roles Resources, Writing – review & editing

    Affiliation Laboratoire Evolution, Génomes et Spéciation (LEGS), CNRS, France

  • Carlos R. Vilela,

    Roles Data curation, Resources, Writing – review & editing

    Affiliation Departamento de Genética e Biologia Evolutiva, Universidade de São Paulo, São Paulo, SP, Brazil

  • Antonio Bernardo Carvalho

    Roles Conceptualization, Formal analysis, Funding acquisition, Investigation, Methodology, Resources, Supervision, Writing – original draft, Writing – review & editing

    bernardo1963@gmail.com, bernardo1963@ufrj.br

    Affiliation Departamento de Genética, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil

Abstract

Y chromosomes are widely believed to evolve from a normal autosome through a process of massive gene loss (with preservation of some male genes), shaped by sex-antagonistic selection and complemented by occasional gains of male-related genes. The net result of these processes is a male-specialized chromosome. This might be expected to be an irreversible process, but it was found in 2005 that the Drosophila pseudoobscura Y chromosome was incorporated into an autosome. Y chromosome incorporations have important consequences: a formerly male-restricted chromosome reverts to autosomal inheritance, and the species may shift from an XY/XX to X0/XX sex-chromosome system. In order to assess the frequency and causes of this phenomenon we searched for Y chromosome incorporations in 400 species from Drosophila and related genera. We found one additional large scale event of Y chromosome incorporation, affecting the whole montium subgroup (40 species in our sample); overall 13% of the sampled species (52/400) have Y incorporations. While previous data indicated that after the Y incorporation the ancestral Y disappeared as a free chromosome, the much larger data set analyzed here indicates that a copy of the Y survived as a free chromosome both in montium and pseudoobscura species, and that the current Y of the pseudoobscura lineage results from a fusion between this free Y and the neoY. The 400 species sample also showed that the previously suggested causal connection between X-autosome fusions and Y incorporations is, at best, weak: the new case of Y incorporation (montium) does not have X-autosome fusion, whereas nine independent cases of X-autosome fusions were not followed by Y incorporations. Y incorporation is an underappreciated mechanism affecting Y chromosome evolution; our results show that at least in Drosophila it plays a relevant role and highlight the need of similar studies in other groups.

Author summary

In contrast to other chromosomes (X and autosomes), which are present in males and females, Y chromosomes spend all time in males. Hence it is not surprising that along evolution they became male specialized, e.g., containing a disproportionate amount of male-fertility genes. Interestingly it was found in 2005 that in Drosophila pseudoobscura the Y chromosome reverted to "male-female existence", being incorporated into an autosome. These "Y chromosome incorporations" have important consequences on sex-chromosome evolution, and allow the study of the evolutionary forces that shaped Y chromosomes as they act backwards. As D. pseudoobscura was the second Drosophila species investigated in this respect, it is likely that other cases exist, and that perhaps it is a common phenomenon. In order to answer this question we studied 400 Drosophila species. We found one additional case of Y incorporation, which occurred in the ancestor of Drosophila montium, and currently affects a large number of species; overall 13% of the species we sampled (52/400) have Y incorporations. We also found that a previously suggested cause of Y incorporations (X-autosome fusions) is not a general explanation. Our results show that in Drosophila Y incorporations play a relevant role and highlight the need of similar studies in other groups.

Introduction

Sex-chromosomes are believed to originate from an ordinary pair of autosomes, after one of them acquires a strong male (or female) determining gene ("M"), becoming a proto-Y (or proto-W; for the sake of simplicity we will refer to both W and Y chromosomes as "Y chromosome"). Natural selection would then favor the accumulation of sex-antagonistic alleles (male-beneficial / female-detrimental) in the vicinity of the M locus, and the gradual suppression of recombination between the proto-X and proto-Y. The absence of recombination ultimately leads to massive gene loss in the proto-Y, with preservation of a few genes; a few male-related genes are also usually acquired from the other chromosomes. The proto-Y becomes a "mature" Y chromosome, containing the male sex-determining gene, some other male related genes (e.g., male-fertility factors), a few relic house-keeping genes (most of them with homologs in the X), and a large amount of repetitive DNA [13]. This "canonical route" is best illustrated and empirically supported by mammalian Y chromosomes: the human Y encodes only 27 proteins (among them, the product of the master sex-determining gene SRY) and 18 of them are shared with the X, whereas the X has a gene content compatible with its size (~1100 genes; [4,5]). Hence the mammalian Y is largely an impoverished ("degenerated") version of the X chromosome [5,6]. This process seems to have occurred independently in many other groups, such as birds, fishes, and plants [1,7,8]. It must be noted, however, that the most informative data—gene content of the Y—is well known in very few species, even after genome sequencing: as with other repeat-rich regions, Y chromosomes are very hard to assemble, requiring special methods that are time consuming and/or expensive [5,916]. When such data became available in Drosophila it revealed striking discrepancies with the canonical route (reviewed in ref [17]).

Given its male-restricted condition it is not surprising that the gene content of Y chromosomes usually is heavily male-biased [5,11]. Furthermore, for Y chromosomes that evolved through the canonical model, accumulation of male-beneficial alleles lies at the heart of their origin. Hence cases in which Y chromosomes revert to autosomal inheritance are particularly interesting because (i) they offer the unique opportunity to study the forces that shaped the Y chromosome (lack of recombination; male-restricted status, reduced effective population size) after they disappear or are reversed; (ii) they allow the study of the origin of Y chromosomes (below). That such reversals occur is hinted by the observation of turnover of sex-chromosomes (e.g., refs. [1820]). One of the best known cases was discovered in 2005: in the D. pseudoobscura lineage the ancestral Y chromosome became part of an autosome, and was replaced by a Y chromosome with unknown origin; the formerly Y-linked genes now have autosomal inheritance [2123]. As a consequence, the current D. pseudoobscura Y chromosome shares no genes with the ancestral Drosophila Y (which is present in D. melanogaster and most species), despite their similarities (both pair with the X, are heterochromatic, and are essential for male fertility). Following Ellison [24], we use here the term "Y incorporation" (meaning "Y chromosome incorporation into an autosome or the X") to distinguish this phenomenon from the usual Y-autosome fusions, in which the ancestral Y keeps its sex-chromosome and male-restricted state, and the fused autosome become a neoY [2527]. Y incorporations and replacements have a direct bearing on the origin and evolution of these chromosomes, and are the main subject of the present paper. We aim here to answer two questions: (i) How frequent are Y incorporations, as happened in the D. pseudoobscura lineage? (ii) It has been suggested that in D. pseudoobscura lineage the Y incorporation was an adaptive response to a (presumably) previous event of X-autosome fusion, which created X-Y segregation problems [21]. Is this a robust explanation? The sample of 12 Drosophila species analyzed before [28,29] is insufficient to answer these questions. Here we report a search for Y incorporations and individual gene movements in 400 species of Drosophila and related genera.

Results

As we are searching for potentially rare events we used the largest possible sample (400 species;Table A in S1 File). All these species are phylogenetically nested within the Drosophila genus, although some of them formally belong to other genera (e.g., Zaprionus; Methods); for the sake of simplicity we will refer to all of them as Drosophila. Genome sequencing at this scale is too expensive, so in order to detect Y incorporations we used degenerate PCR, targeting nine known Y-linked genes (Methods). As shown in Fig 1, in Y incorporations all tested genes shift from male-specific PCR amplification (Y-linkage) to male and female amplification (X or autosomal linkage; hereafter, "X/A" linkage; [21]). Besides this large-scale events, there are also individual gene losses from the Y, which in Drosophila occur by two mechanisms: (i) more commonly, the formerly Y-linked gene moves to an autosome (or the X), in which case a single gene shifts to male and female amplification ("Y chromosome gene transfer"); (ii) the formerly Y-linked gene is lost from the genome, in which case PCR systematically fails in both sexes ("genomic gene loss") [28]. The PCR approach has some limitations, but they do not affect the detection of Y incorporations (Supporting Information). It was very reliable: 94% (3378/3593) of the degenerate PCRs worked (Table B in S1 File), and we readily detected the previously known Y incorporation of the D. pseudoobscura lineage (Table A in S1 File). These raw PCR results (exemplified in Fig 2A for the ORY gene) must be placed into a phylogenetic context in order to identify the independent events (Fig 2B). An analogous procedure was applied to the nine genes in the 400 species (Table A in S1 File). Overall, in the whole dataset we found 21 independent events in which one or more formerly Y-linked genes shifted from male-only to male-female PCR amplification (Table C in S1 File).

thumbnail
Fig 1. Primary data for detection of Y incorporations and individual gene movements.

A) Degenerated PCR of four genes (kl-2, kl-3, Ppr-Y and ORY) in males and females of D. nannoptera and D. pseudoobscura. These four genes are part of the ancestral Y chromosome of the Drosophila genus. Note that in D. nannoptera the ORY gene shifted to amplification on both sexes, indicating an individual gene movement, whereas in D. pseudoobscura all genes were affected, indicating a Y incorporation event. Analogous experiments were performed for nine genes in 400 species. kl-2 primers: kl2_KVME_F1 / kl2_QMQE_R1 (860 bp); kl-3: kl3_DKMD_F / kl3_EMQD_R (1140 bp); Ppr-Y: PprY_FVEH_ns_F1 / PprY_MHGE_R1 (380 bp); ORY: ORY_YKNI_F1 / ORY_IEKE_R1 (500 bp).

https://doi.org/10.1371/journal.pgen.1007770.g001

thumbnail
Fig 2. Phylogenetic interpretation of gene movements.

A) Degenerated PCR of the ORY gene shows that it is autosomal or X-linked in three Zaprionus species and in D. nannoptera, and Y-linked in the remaining species. Species in the order of appearance: D. melanogaster, D. virilis, D. acanthoptera, D. nannoptera, D. immigrans, Z. africanus, Z. bogoriensis, and Z. davidi. B) Given the known phylogeny of the species [66,70] the data is best explained by two individual gene movements (blue dots) of the ORY gene from the Y chromosome to an autosome or the X ("A/X"), ORY being ancestrally Y-linked; the alternative hypothesis of ORY being ancestrally A/X implies more than 10 movements to the Y. Note also that the A/X status of the three Zaprionus species traces back to only one independent event, at the root of the genus (marked with the blue dot). For the sake of clarity we omitted the Y chromosome incorporation events. PCR primers: ORY_YKNI_F1 / ORY_IEKE_R1 (500 bp).

https://doi.org/10.1371/journal.pgen.1007770.g002

Events affecting one Y-linked gene

A total of 17 Y chromosome gene loss events affected only one gene; 16 of them were Y chromosome gene transfers, and one was a genomic gene loss (the Ppr-Y gene in Hawaiian Drosophila; ref [28]). Individual gene losses are a well known phenomenon in Y chromosome evolution in Drosophila and other organisms [6,28,30]; in the 400 species dataset our best estimate of the rate of gene loss is 0.000985 gene lost / gene / Myr (95% CI: 0.000525–0.001685), which is not significantly different (P > 0.5; two-tailed exact test for the ratio of two Poisson means; ref [31]) from a previous estimate based on a much smaller sample (0.001026 gene lost / gene / Myr, 95% CI: 0.000124–0.003707; [14,28]; Supporting Information). We also detected some gene gains by the Y (Supporting Information), but since they cannot be unbiasedly detected by our PCR approach they will not be further discussed.

Besides these 17 events that affected one gene, we found one event simultaneously affecting two genes: in the ancestor of the obscura group the PRY and JY-alpha genes most likely moved from the Y chromosome to the X in a single event, because the two genes are side-by-side on the D. pseudoobscura and D. obscura X-chromosomes [32,33]; it would be unlikely that two independent events moved these formerly Y-linked genes to another chromosome in the same position (Fig A in S1 File). Given this, we conservatively counted the PRY / JY-alpha movements as a single gene loss (counting as two losses, or excluding them, does not change any conclusion).

Events affecting all Y-linked genes

In three events all (or nearly all) genes formerly present in the Y chromosome became present in females. (Table C in S1 File). These are the most interesting cases, since they strongly indicate Y incorporations. The three cases are described below.

The case of the D. pseudoobscura lineage is known [21], and indeed was the main motivation for the present study.

The second event affected one species of the repleta group, D. limensis. However, examination of the mitotic chromosomes strongly indicates that D. limensis does not represent a naturally occurring case of Y incorporation, but rather a laboratory artifact which happened along the decades in which the D. limensis stock has been maintained in culture (Supporting Information).

The third event affected nearly all tested species of the montium subgroup (melanogaster group), and hence presumably occurred at the common ancestor of this large taxon (~100 described species;[34]). The event clearly affected the whole Y chromosome, since in many species all genes that are Y-linked in the montium ancestor became present in females (Fig 3). Many montium species have been investigated cytogenetically; all are XY/XX and have an X chromosome with a single arm [35,36]. These observations make the montium case very interesting, because the absence of a two-armed X implies that the current Y chromosome cannot be a neoY chromosome, as has been suggested for D. pseudoobscura [21,22]. As detailed in the next section, the case of the montium subgroup proved to be even more interesting.

thumbnail
Fig 3. Phylogeny and gene content of the Y chromosome in the montium subgroup.

“M” means amplification only in males (i.e., Y-linkage), whereas “MF” means amplification in both sexes (autosomal or X-linkage). The ancestral state of the montium subgroup is inferred to be similar to D. ananassae and D. rhopaloa (i.e., all 11 genes Y-linked; orange branches). Note that all species of the “clade kikkawai” (red branches) have kl-2 and WDY genes only in males, while the species of the “clade auraria” (blue branches) all ancestrally Y-linked genes are present in both sexes. The basal clade “parvula-kanapiae” (green branches) shared kl-5 and PRY genes only in males. Phylogeny based on Amyrel gene sequences (ML, bootstrap 1,000 replicates).

https://doi.org/10.1371/journal.pgen.1007770.g003

Y incorporation in the montium subgroup

As we sampled more montium species and genes we obtained an initially puzzling result: most of the ancestrally Y-linked genes sporadically "re-acquire" their Y-linkage along the phylogeny. For example, ORY is Y-linked in one of the 40 tested species; PPr-Y in three species, and so forth. We attributed these initial cases to independent re-acquisition of male genes by the Y chromosome, but in order to clarify the phenomenon we made an effort to increase the number of montium species (by asking colleagues and stock centers) and tested genes (by adding non-degenerate primers based on the D. kikkawai genome sequence [37]; Supporting Information). We also made a molecular phylogeny of the montium species we studied, using Amyrel gene sequences produced by DaLage and co-workers [34] and by ourselves (accession numbers MH707309- MH707325). The resulting phylogeny is similar to previous studies [38,39], which do not contain all species we tested. Fig 3 integrates the phylogeny and Y-linkage data. Three main phylogenetic clades of montium species appear, and each has a characteristic Y-linked gene content. In the auraria clade (eight species in our sample) all formerly Y-linked genes have A/X linkage, supporting the pattern of "complete Y incorporation" seen before in D. pseudoobscura [21]. In the kikkawai clade (24 species) the WDY and kl-2 genes are always Y-linked; besides them, several other genes are Y-linked in a few scattered species. The third phylogenetic clade, composed by D. kanapiae and D. parvula, has kl-5 and PRY (as well as several other genes) on the Y chromosome. The gene content of the Y chromosomes seems to have no correlation with their morphologies, which are quite variable among montium species (Table D in S1 File). Similarly, D. lini and D. kikkawai have been shown to carry B chromosomes [40,41], which are non-essential supernumerary chromosomes that have been suggested as a possible origin of the Drosophila Y chromosome [17,42]. Note that the Y-linked gene content of these two species (Fig 3) is similar to other species of the "kikkawai clade" (which do not carry B chromosomes); this suggests that the two phenomena (changes in Y-linked gene content and presence of B chromosomes) are unrelated, although we must note that we have not examined the specific strains used in the present work for the presence of B chromosomes. Finally, using the D. melanogaster Y chromosome cytogenetic map as a proxy for the montium species, we found no association between the changes in linkage shown in Fig 3 and the cytogenetic position of the genes (Fig B in S1 File).

We can summarize the pattern we found in montium species as follows: previously Y-linked genes seem to "re-appear" in the Y chromosome, closely related species tending to share the same genes on the Y (Fig 3). Before further investigation of this pattern it is worth to re-examine the D. pseudoobscura case.

Y incorporation in the D. pseudoobscura lineage revisited

Previous data strongly suggests that the Y incorporation in the D. pseudoobscura lineage (i.e., the pseudoobscura and affinis subgroups) is complete and irreversible: in all tested species (D. pseudoobscura, D. persimilis, D. miranda, D. azteca, and D. affinis) all formerly Y-linked genes (kl-2, kl-3, PPr-Y, ORY, and ARY) acquired autosomal linkage [21]; mapping and sequencing experiments showed that they moved to the small "dot chromosome" [22], amazingly remaining side-by-side [21,23]. However, given the findings on the montium species, one may also think that the pseudoobscura lineage is similar to what we would had observed in montium if we had sampled only the auraria clade (Fig 3). In other words, it is possible that Carvalho and Clark [21] have not sampled enough species to detect the "re-appearance of Y-linkage" in the pseudoobscura and affinis subgroups. Hence we made an effort to obtain "difficult" species of these subgroups (D. helvetica, D. athabasca, D. narragansett, and D. lowei; we are indebted to our colleagues listed in Table A in S1 File, who sent us these and other samples). Indeed, we found in two species of the D. pseudoobscura lineage the same "re-appearance of Y-linkage" phenomenon we observed in the montium subgroup (Fig 4).

thumbnail
Fig 4. Phylogeny and gene content of the Y chromosome in the obscura group.

“M” means amplification only in males (i.e., Y-linkage), whereas “MF” means amplification in both sexes (autosomal or X-linkage). The ancestral state is all genes in the Y chromosome, and is represented by D. ananassae (melanogaster group) and D. bifasciata (obscura subgroup). All remaining species belong to the pseudoobscura and affinis subgroups, and carry the Y incorporation described in ref [21]. Note that three genes (ARY, kl-2 and ORY) seem to have re-acquired Y-linkage in a few species (see Results for the interpretation).Only genes that are Y-linked in the ancestor of the obscura group are shown. The D. athabasca strain came from Rochester, NY, and hence most likely belongs Eastern A subspecies [71]. Phylogeny taken from ref [72].

https://doi.org/10.1371/journal.pgen.1007770.g004

Why genes seem to return to the Y chromosome after Y incorporations?

As we commented before we initially attributed these sporadic re-acquisitions of Y-linkage to the normal process of acquisition of male genes by the Drosophila Y chromosome. However, in the whole set of montium species, nine out 11 genes are Y-linked in at least one species (Fig 3). This seems too frequent to be explained by independent gene gains in a rather short time span (the montium / melanogaster split occurred 41 Myr ago; [43]), unless the gene gain rate by the Y chromosome is very high in the montium subgroup. We formally evaluated this hypothesis as follows. Suppose that the montium data is explained by a high rate of A/X to Y gene movements. This being true, the montium Y should had acquired other male genes besides the nine mentioned above. In order to test this, we took a sample of male genes that are autosomal in the ancestor of the montium subgroup, and that can be acquired by the Y (because they are Y-linked in other Drosophila species), and check if they moved to the Y of montium. We tested four such genes (Y-linked in D. virilis: CG11719, CG2964; Y-linked in D. willistoni: CG18155, CG14339; ref [14]), and in all four genes there is no case of A/X to Y movement in the whole set of 40 montium species. The difference between 9 genes out 11 and 0 genes out 4 is statistically significant (P = 0.011, Fisher exact test; Table 1; see also Supporting Information, section "Statistical test of the re-acquisition of Y-linked genes"). Hence the re-appearance of Y-linkage cannot be explained by a high gene gain rate by the Y chromosome, because only genes that formerly were Y-linked are affected.

thumbnail
Table 1. Re-appearance of Y-linked genes in the montium subgroup.

https://doi.org/10.1371/journal.pgen.1007770.t001

A similar observation holds for the D. pseudoobscura lineage: again there is statistically significant evidence that only formerly Y-linked genes re-acquired Y-linkage (P = 0.035, Fisher exact test; Table E in S1 File).

Using the montium species as an example, the most parsimonious hypothesis to explain the pattern found in both cases is that a free copy of the Y remained in the genome when the Y chromosome was incorporated into an autosome (or the X) in the ancestor of the montium subgroup (Fig 5). Each formerly Y-linked gene would then have two non-allelic copies: the Y-linked and the A/X-linked one. The outcome of this genetic redundancy would have been a more or less random loss of either the free Y copy, or the A/X-linked one [44]. This process would have occurred along with the diversification of the montium subgroup, generating the current pattern in which previously Y-linked genes seem to "re-appear" in the Y chromosome, closely related species tending to share the same genes on the Y (Fig 3). As shown in Fig 5, by chance some species would had lost all Y-linked copies (e.g., the auraria clade), whereas others would had lost most of the autosomal copies (e.g., D. nikananu). The same explanation, called hereafter the "duplicated Y" hypothesis, applies to the D. pseudoobscura lineage. We will return to this point in the Discussion.

thumbnail
Fig 5. The duplicated Y hypothesis as an explanation for the re-appearance of Y-linkage after Y incorporations.

The ancestor of the montium subgroup was similar to D. melanogaster (top). Then an incorporation of the Y chromosome into an autosome would have generated an intermediate stage with two genetically redundant copies of each formerly-linked gene (middle). Random losses of one redundant copy of each gene would result in species like D. auraria, which is devoid of any Y-linked gene, and D. parvula and D. kanapiae, in which some genes seem to have re-acquired Y-linkage (bottom). This hypothesis explains why only formerly Y-linked genes seem to have moved to the Y chromosome (Table 1), and why different montium species may carry different subsets of these Y-linked genes (Fig 3); it also explains the D. pseudoobscura lineage data (Table E in S1 File and Fig 4).

https://doi.org/10.1371/journal.pgen.1007770.g005

Discussion

Incorporation of the Y chromosome into other chromosomes is a very interesting phenomenon with important consequences: (i) formerly male-restricted genes revert to autosomal inheritance and become present in females; (ii) the species may shift from a XY/XX to X0/XX sex-chromosome system, or a new Y chromosome may appear; (iii) the evolutionary forces and conditions that shaped Y chromosome evolution (lack of recombination; male-restricted status, reduced effective population size) disappear or are reversed. These changes have a direct bearing on important aspects of sex-chromosome evolution theory, such as the role of sex-antagonistic genes on the evolution of Y chromosomes [45,46], and the origin of Y chromosome itself.

Y incorporations occurred at least twice in the Drosophila clade

The present work was motivated by the finding that in the second sequenced Drosophila species (D. pseudoobscura; ref [47]) there was a Y incorporation [21]. This strongly suggested that other cases exist, and that perhaps it is a common phenomenon. The answer to these two questions are respectively "yes" and a qualified "no": while studying 400 species (6,529 Myr of evolution), we indeed found one additional case of Y incorporation (the montium subgroup); on the whole, 13% of our sample (52 species out of 400) has a Y incorporation; the rate is 0.000893 Y incorporations / Myr (Supporting Information). This rate is similar to the rate of individual gene loss per gene calculated above for the same 400 species, i.e., a given Drosophila Y-linked gene has roughly the same chance of being lost individually, or by a Y chromosome incorporation event. Y chromosome incorporations, as individual gene losses in the Drosophila Y, are rather rare events; however, if we consider a broader time scale (e.g., the Diptera order) they may be frequent enough to explain, for example, the lack of homology between the Y chromosome of Drosophila and more distant Diptera such as Tephritidae and mosquitoes [28,48]. Analogously, the identity of the X chromosome (Muller element A) is conserved across the Drosophilidae family, but not in Diptera [49].

Are there additional cases of Y incorporations in the phylogenetic clade studied here? Probably yes. For example, although not represented in our sample, there are Samoaia and Hirtodrosophila species with X0/XX sex determination [24,50], in which presumably the ancestral Y (and its genes) was incorporated into another chromosome (the Samoaia and Hirtodrosophila species we sampled have XY males, and the normal gene content in the Y; Table A in S1 File). This is exactly what we found in X0 Chymomyza species (e.g., C. procnemis; ref [51]): the ancestrally Y-linked genes are present in males and females (unpublished).

It is also interesting to consider the relevance of Y incorporations outside Drosophila. The discovery of Y incorporations usually requires genome sequencing and a detailed investigation of the Y-linked genes, so they easily went unnoticed, as shown by the cases of D. pseudoobscura (a species studied since the 1920's) and D. kikkawai (which Y incorporation was not detected, despite the species being sequenced a few years ago; [37]); this suggests that there may be cases even in fairly well known species. On the other hand, species with a sex-determining Y chromosome (e.g., many vertebrates and Diptera) most likely are less prone to Y incorporations because such events are expected to disrupt the sex-ratio. However, note that in many such groups there had been cases of turnover of the sex-determination system (e.g., [5254]), which most likely had similar effects on the sex-ratio. Hence, the relevance of Y incorporations outside Drosophila remains to be determined, and it would be desirable to perform similar studies in other groups.

Natural selection and Y incorporations

Our study offers some clues on the possible fitness advantages of Y chromosome incorporations. Carvalho and Clark [21] found that the Y incorporation is present in the closely related affinis and pseudoobscura subgroups (which are known to carry an X-A fusion), and is absent in the more distantly related obscura subgroup (which does not carry the X-A fusion), and suggested that the Y incorporation was an adaptive response to the X-A fusion. Namely, X-A fusions are known to cause meiotic problems because three (instead of two) centromeres must correctly pair and segregate in males: the X (fused with one member of the autosomal pair), the ancestral Y, and the other member of the autosomal pair (called "neoY"); incorporation of the ancestral Y into another autosome would solve this problem (see Fig 3 of ref [21]). This is an attractive hypothesis because it seems to explain very well the puzzling phenomenon found in the D. pseudoobscura lineage. However two lines of evidence found here suggest that the causal link between X-A fusions and Y chromosome incorporations is, at best, weak. First, there are 35 species (nine independent events) with an X-A fusion similar to D. pseudoobscura in our sample of 400 species, and all carry the ancestral Drosophila Y (Table F in S1 File). Therefore Y incorporation is not an obligatory consequence of X-A fusions. Second, in the other case of Y incorporation we found (montium subgroup) the X chromosome is not fused to any autosome (it has a single arm). Therefore Y incorporations can occur in the absence of X-A fusions. Thus, at this point, the relationship between X-A fusions and Y incorporation is unclear. Related to this, we do not know if Y chromosome incorporations are selection-driven events, as proposed by Carvalho and Clark [21] or, alternatively, mutations fixed by genetic drift; it is also possible that some were selection-driven events (perhaps D. pseudoobscura), whereas others were fixed by drift (perhaps in species devoid of X-A fusions). It is interesting to note that in other chromosomal mutations such as inversions and translocations the relative roles of selection and drift are also unclear [26,27,55].

Finally, given the data shown in Fig 4 and Table E in S1 File, it seems likely that the current Y chromosome of the D. pseudoobscura lineage actually is an impoverished copy of the ancestral Y chromosome. If this is true, then the explanation offered by Carvalho and Clark [21] for the Y incorporation would be less compelling (for there would had been a free Y chromosome all the time), but not invalidated: one of the fitness costs caused by the X-A fusion is the generation of some sterile X0 sons; the Y incorporation ameliorates this because it would rescue the fertility of these sons.

Alternative hypothesis for the re-appearance of Y-linkage in the montium and pseudoobscura lineages

The duplicated Y hypothesis (Fig 5) is bold, but as far as we can see it is the best explanation for the data. The only alternative hypothesis we could think to explain the pattern shown in Table 1 and Table E in S1 File is that the free Y really disappeared (eventually being replaced by something else) or lost all its genes after its incorporation into an autosome or the X, but there was a high fitness cost of the new location of these formerly Y-linked genes. This would have "accelerated" the movement to the Y (i.e., increased the fixation probability) only for this set of genes, which would explain the data shown in Table 1 and Table E in S1 File. However, if there is such strong selection favoring the Y-linkage of these particular genes, it is difficult to explain how the Y would get lost (or emptied) at first place.

Perhaps the main difficulty of the duplicated Y hypothesis is that it requires that several genes retained their two copies (A/X and Y-linked) for rather long periods across the phylogeny. For example, the re-appearance of Y-linkage for several genes in D. nikananu (kl-5, PRY, ORY, ARY), coupled with their A/X status in the sister species D. diplacantha, would require that the two copies of these genes had persisted since the Y incorporation (at least ~19 Myr ago) until the D. nikananu / D. diplacantha divergence (~5.9 Myr ago; Fig 3 and Fig C in S1 File). A similar reasoning applies to several other genes and sections of the phylogeny. This retention of duplicate genes for ~13 Myr seems too long, since the half-life of duplicated genes in Drosophila has been estimated as 0.66 Myr (ref [56]) and 3.2 Myr (ref [44]). However, there is a major factor that may explain the discrepancy between these half-life estimates and ~13 Myr: while the former were calculated from single gene (or small segment) duplications, in the later a whole chromosome was duplicated. So, in the Y incorporations the genes were duplicated along with all long-range regulatory regions and chromatin state signals, which most likely favors a long survival. In this sense Y incorporations resemble polyploidizations, which are known to have an unexpectedly long preservation of duplicated genes (tens to hundreds of millions of years; ref [57], pp. 207–208, and references cited therein).

Possible tests of the duplicated Y hypothesis

There are several ways to test this hypothesis. First, as it implies that the locations of the "re-acquired" Y-linked genes actually are ancestral, it predicts that gene order in the Y chromosome of species of the montium subgroup (e.g., D. nikananu and D. kanapiae; Fig 3) should be conserved among themselves and perhaps with D. melanogaster [58,59], although we should keep in mind that nothing is known about the rate of rearrangements in the Drosophila Y. Such test must be done by fluorescent in situ hybridization (FISH), or by a long-read assembly [23], since Sanger or Illumina assemblies are too fragmented to provide synteny information of repetitive regions. Second, since two gene copies (Y and autosomal/X-linked) seem to have co-existed for a long time, one might expect to see both in some species. To date D. kikkawai and D. serrata are the only montium species that has been sequenced [37,60], and we found no signs of two gene copies either in the assembled scaffolds or the raw traces (not shown), but both assemblies are not good for this purpose (see section "Genomic effects of the Y incorporation in montium subgroup species"). We are now obtaining improved assemblies of both species to search for duplicated genes. Finally, full identification of the Y-linked genes (e.g., ref [14]) in the montium subgroup and in D. pseudoobscura (below) from may help to elucidate the origin of their current Y chromosomes. We are now pursuing some of these approaches.

On the origin of the D. pseudoobscura Y

The results described in this work shed light on the origin of the current Y chromosome of the D. pseudoobscura lineage, a subject that has been somewhat controversial. Carvalho and Clark [21] suggested that it may be the neoY (i.e., the remnant of the homolog of the Muller-D autosome that got fused with the X), but noted that direct evidence (e.g., a concentration of Muller D-derived genes in the Y) was lacking, and that other origins are possible. Other authors assumed that the current Y is a neoY. As we noted above, the data presented in this work strongly suggest that the current Y of this lineage is a very impoverished ancestral Y. However, ongoing work in our lab identified four functional protein coding genes in the D. pseudoobscura Y; three of them have orthologs in D. melanogaster which are located in the Muller D (CG6661, CG6845, and CG32181; the fourth gene, GA27172, does not have a clear ortholog; [61]). So, as shown in Fig 6, the current Y of D. pseudoobscura lineage seems to be both a remnant of the ancestral Y (as indicated by the "re-appearance" of Y-linkage) and a remnant of the neoY (as indicated by the three Muller-D derived genes). Actually, this dual nature of the D. pseudoobscura Y was suggested by M.J.D. White more than 40 years ago: "Most probably, this loss of the Y2 [the Muller-D derived neoY] took place through its fusion with the original Y (= Y1) and subsequent heterochromatinization." (ref [26], p. 350). White was not aware of the incorporation of the ancestral Y into an autosome [2123], which is a major phenomenon, but other than this his suggestion probably is correct.

thumbnail
Fig 6. Updated model for the origin of D. pseudoobscura sex chromosomes.

Autosomes are shown as solid bars (shown three pairs, A1, A2 and A3), sex chromosomes as striped bars, and centromeres as open circles. (A) The ancestral state configuration is typical for Drosophila (e.g., obscura subgroup species). (B) A centric fusion between the X and one autosome (Muller element D, corresponding to chromosome 3L in D. melanogaster; [73]) generated an X Y1 Y2 sex-determination system, which was transient in the D. pseudoobscura lineage, but exists in other species [26]. (C) A Y incorporation event placed the ancestral Y (Y1) within another autosome (Muller element F, corresponding to chromosome 4 in D. melanogaster; [2123]). The present work strongly suggests that a copy of the ancestral Y (Y1) survived also as a free chromosome and fused with the neoY (Fig 4 and Table E in S1 File), originating the current Y chromosome of the D. pseudoobscura lineage. The order of the events represented by the two large arrows is hypothetical. Also, the Y incorporation and the free Y—neoY fusion events (represented here by a single arrow) probably were two separate events. Modified from reference [21].

https://doi.org/10.1371/journal.pgen.1007770.g006

Genomic effects of the Y incorporation in montium subgroup species

Y incorporations offer the unique opportunity to study the effect of the forces that shaped the Y chromosome (lack of recombination; male-restricted status; reduced effective population size) as they disappear or are reversed. We will focus here on two characteristics of Drosophila Y-linked genes: intron size and intergenic distance. Drosophila Y-linked genes contain some Mbp-sized introns, filled with repetitive DNA [59,62]. These huge introns cause assembly breaks, so multi-exon genes are never assembled in a single scaffold; intergenic sequences are also large and repeat-rich, so we seldom observe two genes in the same scaffold [11,28]. Interestingly, in the incorporated Y of D. pseudoobscura the intron sizes became fairly normal and intergenic distances were reduced [21,23]. As a consequence, even in a Sanger assembly large genes (kl-2, kl-3) were fully assembled, and most genes are present in scaffolds containing more than one gene [21]. So it seems that after being incorporated into an autosome the former Y is acquiring genomic characteristics of euchromatic regions, i.e., we would be observing backwards some of the processes of Drosophila Y evolution. Larracuente and Clark [63] suggested that positive selection to reduce intron sizes played a role in this process: the incorporated Y chromosome of D. pseudoobscura shows a strong reduction of genetic variability (which is compatible with recurrent selective sweeps), in the absence of detectable positive selection in the coding sequence of the genes (which suggest that non-coding regions were the target of selection).

The discovery of the case of Y incorporation in the montium subgroup allows us to further investigate this phenomenon. As shown in Table G in S1 File (columns 5–6), the incorporated Y region of both montium species with sequenced genomes (D. kikkawai and D. serrata [37,60]) seems to be much more similar to free Y chromosomes (e.g., D. melanogaster) than to the incorporated Y of D. pseudoobscura: the genes are heavily fragmented (a sign of large, repeat-rich introns) and in only one case is there more than one gene in the same scaffold (a sign of large, repeat-rich intergenic spaces).

If confirmed, the above findings would be particularly interesting because the Y incorporation in montium, which occurred 19 to 41 Mya (Fig C in S1 File), seems to be older than the D. pseudoobscura event (12.7 to 20.8 Mya; ref [64]), and hence the "euchromatinization" would be proceeding with much slower speed in montium species, or not occurring at all. However, there is a big caveat: the genome assemblies of both montium species are much more fragmented and/or have assembly issues that make them not directly comparable to D. pseudoobscura (Table G in S1 File, columns 7–8). Namely, D. kikkawai assembly used short reads (Illumina and Roche/454) which is not good for repetitive regions [65], whereas the D. serrata PacBio assembly is also fairly fragmented and with other assembly problems in the incorporated Y genes (e.g., out-of-frame indels and missing exonic sequences; Table G in S1 File). The D. pseudoobscura assemblies, which were based in Sanger or in PacBio [21,23], do not have these problems. It will be very interesting to study the incorporated Y region in improved assemblies of montium species.

Concluding remarks

The finding of a Y incorporation in the second Drosophila species that was sequenced (D. pseudoobscura; refs [21,22]) strongly suggested that there are other cases, and that this may be a common phenomenon. These questions can only be answered empirically. We did this with a sample of 400 Drosophila species, and found one additional case (the montium subgroup), besides the previously known event in the D. pseudoobscura lineage. These two events affect 13% of the sampled species (52/400); they happened in the last ~60Myr, and projected into larger time scale, may explain the complete lack of Y chromosome homology between the Drosophila and more distant Diptera such as mosquitoes and Tephritidae [48]. In both the montium and pseudoobscura lineages the Y incorporation resulted in XY/XX species in which the Y seems to be derived from the ancestral Y chromosome; ongoing work on the D. pseudoobscura lineage suggests that its Y also contain the remnants of the neoY. Finally, the data presented here show that the formerly suggested adaptive explanation for Y incorporations [21] is not general, and hence this remains an open question. These findings, the discoveries that X chromosomes are also replaced [49], and that the evolution of the Drosophila Y is dominated by gene gains instead of gene losses [14,28], show that sex-chromosome evolution in Diptera is a dynamic and complex process, and that there is a lot more to be learned about it.

Methods

Species studied and DNA extraction

The 400 Drosophilid species (Table A in S1 File) belong to the following formally recognized genera (subgenera): Dettopsomyia, Drosophila (Dorsilopha, Drosophila, Phloridosa, Siphlodora, Sophophora), Hirtodrosophila, Mycodrosophila, Samoaia, Scaptomyza (Bunostoma, Parascaptomyza), Zaprionus (Anaprionus, Zaprionus) and Zygothrica, which form a natural group (i.e., a monophyletic clade) with fairly well known phylogenetic relationships [66,67]. We used all samples we could obtain (Table A in S1 File), provided that we got at least one male and one female from the same species. Most samples came from isofemale lines. Unculturable species were represented by copulating pairs (when available) or by wild-caught material, both identified by an experienced Drosophila taxonomist (usually Carlos R. Vilela). DNA was extracted separately from males and females (one to ~ four individuals of each sex); in most cases females may had been inseminated so we used only their thorax to avoid contamination from stored sperm. We used standard phenol-chloroform extraction or the Puregene kit (Qiagen cat # 158667), following the manufacturer's instructions; DNA was quantified with Qubit (Invitrogen cat # Q32857), usually diluted to ~ 1 ng/ul, and frozen at -20 oC until use.

Gene choice and degenerate PCR

The whole set of species was studied with degenerate PCR primers for nine genes: six Y-linked genes that are present in the ancestral Drosophila Y (kl-2, kl-3, ORY, PPr-Y, PRY, JY-alpha), plus the kl-5, WDY and CG11719 genes [14,28,68]. These nine genes were chosen by two criteria: being informative (i.e., Y-linked in a large number of species), and allowing the design of reliable degenerate primers. In the two cases of suspected Y incorporation, additional genes (that do not yield reliable degenerate primers) were tested for Y-linkage with normal PCR primers, designed using a closely related sequenced species: genes Pp1-Y1, Pp1-Y2 and ARY in the montium subgroup (based on the D. kikkawai genome sequence), and the ARY gene in the pseudoobscura lineage (based on D. pseudoobscura). Note that Fig 3 shows only 11 genes (instead of 12) because the CG11719 gene is non-informative in the Sophophora subgenus (it is ancestrally autosomal). Degenerate primers were designed using the Codehop method [69], targeting protein regions that are conserved among the 12 sequenced Drosophila species (Supporting Information). The PCR protocols were deposited in dx.doi.org/10.17504/protocols.io.szyef7w.

Confirmation of PCR results

As contamination of female DNA by male DNA, or spurious PCR amplification might lead to errors in linkage ascertainment (namely, a Y-linked gene be considered autosomal / X-linked), all suspected cases of gene movements were confirmed with at least three PCR experiments, using different strains when available, and also re-testing the closest related species. If they passed these tests we confirmed the result by sequencing the PCR product in females or in both sexes (Supporting Information). Another layer of error control is provided by phylogenetic consistency: among the 21 independent events of linkage changes, 11 affected several or many species, which provides a cross-check. As can be seen in Table A in S1 File, the data are highly consistent. All sequencing was performed at Macrogen (Korea).

Supporting information

S1 Dataset. Excel file used to estimate the loss rate of Y-linked genes.

https://doi.org/10.1371/journal.pgen.1007770.s002

(XLS)

Acknowledgments

We thank all institutions and colleagues who sent the flies used in this work: National Drosophila Species Stock Center (USA), Drosophila Stocks of Ehime University (Japan), L.B. Klaczko, E. Loreto, L. Robe, A. Ludwig, B. Bitner-Mathé, B. Goñi, B. McAllister, M.L. Cariou, M. Polak, J. Jaenike, T. Markow, A. Kopp, A. Clark, M. Watada, D. Bachtrog, M. Noor, and C. Finet. We also thank M. Vibranovski, B. Lemos, A. Yassin, and J. Henikoff for valuable comments and/or help.

References

  1. 1. Bull JJ (1983) Evolution of sex determining mechanisms. Menlo Park, Calif.: Benjamin/Cummings Pub. Co. Advanced Book Program. xx, 316 p.
  2. 2. Charlesworth B, Charlesworth D (2000) The degeneration of Y chromosomes. Philos Trans R Soc Lond B Biol Sci 355: 1563–1572. pmid:11127901
  3. 3. Rice WR (1996) Evolution of the Y sex chromosome in animals. Bioscience 46: 331–343.
  4. 4. Ross MT, Grafham DV, Coffey AJ, Scherer S, McLay K, et al. (2005) The DNA sequence of the human X chromosome. Nature 434: 325–337. pmid:15772651
  5. 5. Skaletsky H, Kuroda-Kawaguchi T, Minx PJ, Cordum HS, Hillier L, et al. (2003) The male-specific region of the human Y chromosome is a mosaic of discrete sequence classes. Nature 423: 825–U822. pmid:12815422
  6. 6. Hughes JF, Rozen S (2012) Genomics and genetics of human and primate Y chromosomes. Annu Rev Genomics Hum Genet 13: 83–108. pmid:22483277
  7. 7. Bachtrog D, Mank JE, Peichel CL, Kirkpatrick M, Otto SP, et al. (2014) Sex determination: why so many ways of doing it? PLoS Biol 12: e1001899. pmid:24983465
  8. 8. Papadopulos AS, Chester M, Ridout K, Filatov DA (2015) Rapid Y degeneration and dosage compensation in plant sex chromosomes. Proc Natl Acad Sci U S A 112: 13021–13026. pmid:26438872
  9. 9. Tilford CA, Kuroda-Kawaguchi T, Skaletsky H, Rozen S, Brown LG, et al. (2001) A physical map of the human Y chromosome. Nature 409: 943–945. pmid:11237016
  10. 10. Hoskins RA, Carlson JW, Kennedy C, Acevedo D, Evans-Holm M, et al. (2007) Sequence finishing and mapping of Drosophila melanogaster heterochromatin. Science 316: 1625–1628. pmid:17569867
  11. 11. Carvalho AB, Lazzaro BP, Clark AG (2000) Y chromosomal fertility factors kl-2 and kl-3 of Drosophila melanogaster encode dynein heavy chain polypeptides. Proc Natl Acad Sci U S A 97: 13239–13244. pmid:11069293
  12. 12. Carvalho AB, Dobo BA, Vibranovski MD, Clark AG (2001) Identification of five new genes on the Y chromosome of Drosophila melanogaster. Proc Natl Acad Sci U S A 98: 13225–13230. pmid:11687639
  13. 13. Carvalho AB, Vibranovski MD, Carlson JW, Celniker SE, Hoskins RA, et al. (2003) Y chromosome and other heterochromatic sequences of the Drosophila melanogaster genome: how far can we go? Genetica 117: 227–237. pmid:12723702
  14. 14. Carvalho AB, Clark AG (2013) Efficient identification of Y chromosome sequences in the human and Drosophila genomes. Genome Res 23: 1894–1907. pmid:23921660
  15. 15. Krzywinski J, Nusskern DR, Kern MK, Besansky NJ (2004) Isolation and characterization of Y chromosome sequences from the African malaria mosquito Anopheles gambiae. Genetics 166: 1291–1302. pmid:15082548
  16. 16. Hall AB, Qi Y, Timoshevskiy V, Sharakhova MV, Sharakhov IV, et al. (2013) Six novel Y chromosome genes in Anopheles mosquitoes discovered by independently sequencing males and females. BMC Genomics 14: 273. pmid:23617698
  17. 17. Carvalho AB, Koerich LB, Clark AG (2009) Origin and evolution of Y chromosomes: Drosophila tales. Trends Genet 25: 270–277. pmid:19443075
  18. 18. Schartl M (2004) Sex chromosome evolution in non-mammalian vertebrates. Curr Opin Genet Dev 14: 634–641. pmid:15531158
  19. 19. Ezaz T, Stiglec R, Veyrunes F, Marshall Graves JA (2006) Relationships between vertebrate ZW and XY sex chromosome systems. Curr Biol 16: R736–743. pmid:16950100
  20. 20. Dufresnes C, Borzee A, Horn A, Stock M, Ostini M, et al. (2015) Sex-Chromosome Homomorphy in Palearctic Tree Frogs Results from Both Turnovers and X-Y Recombination. Mol Biol Evol 32: 2328–2337. pmid:25957317
  21. 21. Carvalho AB, Clark AG (2005) Y chromosome of D. pseudoobscura is not homologous to the ancestral Drosophila Y. Science 307: 108–110. pmid:15528405
  22. 22. Larracuente AM, Noor MAF, Clark AG (2010) Translocation of Y-linked genes to the dot chromosome in Drosophila pseudoobscura. Mol Biol Evol 27: 1612. pmid:20147437
  23. 23. Chang CH, Larracuente AM (2017) Genomic changes following the reversal of a Y chromosome to an autosome in Drosophila pseudoobscura. Evolution 71: 1285–1296. pmid:28322435
  24. 24. Ellison J (1968) A cytological study of some species of the genus Samoaia (Diptera; Drosophilidae). Univ Texas Publs 6818: 421–430.
  25. 25. MacKnight RH (1939) The sex-determining mechanism of Drosophila miranda. Genetics 24: 180–201. pmid:17246918
  26. 26. White MJD (1973) Animal cytology and evolution. Cambridge, Eng.: University Press. viii, 961 p. p.
  27. 27. Pennell MW, Kirkpatrick M, Otto SP, Vamosi JC, Peichel CL, et al. (2015) Y fuse? Sex chromosome fusions in fishes and reptiles. PLoS Genet 11: e1005237. pmid:25993542
  28. 28. Koerich LB, Wang X, Clark AG, Carvalho AB (2008) Low conservation of gene content in the Drosophila Y chromosome. Nature 456: 949–951. pmid:19011613
  29. 29. Drosophila 12 Genomes Consortium (2007) Evolution of genes and genomes on the Drosophila phylogeny. Nature 450: 203–218. pmid:17994087
  30. 30. Dyer KA, White BE, Bray MJ, Piqué DG, Betancourt AJ (2011) Molecular evolution of a Y chromosome to autosome gene duplication in Drosophila. Mol Biol Evol 28: 1293. pmid:21172827
  31. 31. Krishnamoorthy K, Thomson J (2004) A more powerful test for comparing two Poisson means. J Stat Plan Inf 23–35.
  32. 32. dos Santos G, Schroeder AJ, Goodman JL, Strelets VB, Crosby MA, et al. (2015) FlyBase: introduction of the Drosophila melanogaster Release 6 reference genome assembly and large-scale migration of genome annotations. Nucleic Acids Res 43: D690–697. pmid:25398896
  33. 33. Nozawa M, Onizuka K, Fujimi M, Ikeo K, Gojobori T (2016) Accelerated pseudogenization on the neo-X chromosome in Drosophila miranda. Nat Commun 7: 13659. pmid:27897175
  34. 34. DaLage J-L, Kergoat GJ, Maczkowiak F, Silvain J-F, Cariou M-L, et al. (2007) A phylogeny of Drosophilidae using the Amyrel gene: questioning the Drosophila melanogaster species group boundaries J Zool Syst Evol Res 45: 47–63.
  35. 35. Baimai V (1980) Metaphase karyotypes of certain species of the Drosophila montium subgroup. Japan J Genet 55: 165–175.
  36. 36. Shyamala BV, Ranganath HA (1989) Metaphase karyotype differentiation in 8 species of the montium subgroup of Drosophila. Genetica 79: 191–196. pmid:2620831
  37. 37. Chen ZX, Sturgill D, Qu J, Jiang H, Park S, et al. (2014) Comparative validation of the D. melanogaster modENCODE transcriptome annotation. Genome Res 24: 1209–1223. pmid:24985915
  38. 38. Chen CC, Watada M, Miyake H, Katoh TK, Sun Z, et al. (2013) Courtship patterns in the Drosophila montium species subgroup: repeated loss of precopulatory courtship? Zoolog Sci 30: 1056–1062. pmid:24320184
  39. 39. Yassin A, Delaney EK, Reddiex AJ, Seher TD, Bastide H, et al. (2016) The pdm3 locus is a hotspot for recurrent evolution of female-limited color dimorphism in Drosophila. Curr Biol 26: 2412–2422. pmid:27546577
  40. 40. Deng Q, Zeng Q, Qian Y, Li C, Yang Y (2007) Research on the karyotype and evolution of Drosophila melanogaster species group. J Genet Genomics 34: 196–213. pmid:17498617
  41. 41. Sundaran AK, Gupta JP (1994) Accumulation of B-chromosomes in Drosophila kikkawai Burla. Cytobios 80: 211–215. pmid:7774291
  42. 42. Hackstein JHP, Hochstenbach R, HauschteckJungen E, Beukeboom LW (1996) Is the Y chromosome of Drosophila an evolved supernumerary chromosome? Bioessays 18: 317–323. pmid:8967900
  43. 43. Tamura K, Subramanian S, Kumar S (2004) Temporal patterns of fruit fly (Drosophila) evolution revealed by mutation clocks. Mol Biol Evol 21: 36–44. pmid:12949132
  44. 44. Lynch M, Conery JS (2003) The evolutionary demography of duplicate genes. J Struct Funct Genomics 3: 35–44. pmid:12836683
  45. 45. Fisher RA (1930) The genetical theory of natural selection. Oxford,: The Clarendon press. xiv, 272 p. p.
  46. 46. Rice WR (1992) Sexually antagonistic genes: experimental evidence. Science 256: 1436–1439. pmid:1604317
  47. 47. Richards S, Liu Y, Bettencourt BR, Hradecky P, Letovsky S, et al. (2005) Comparative genome sequencing of Drosophila pseudoobscura: chromosomal, gene, and cis-element evolution. Genome Res 15: 1–18. pmid:15632085
  48. 48. Mahajan S, Bachtrog D (2017) Convergent evolution of Y chromosome gene content in flies. Nat Commun 8: 785. pmid:28978907
  49. 49. Vicoso B, Bachtrog D (2013) Reversal of an ancient sex chromosome to an autosome in Drosophila. Nature 499: 332–335. pmid:23792562
  50. 50. Clayton FE (1998) Published karyotypes of the Drosophilidae. DIS 81: 5–125.
  51. 51. Matsuda M (2002) Meiotic chromosome configurations of XY and "XO" type males from four species of the genus Chymomyza (Diptera,Drosophilidae) with a description of the karyotype of C. rufithorax. Cytologia 67: 383–388.
  52. 52. Vicoso B, Bachtrog D (2015) Numerous transitions of sex chromosomes in Diptera. PLoS Biol 13: e1002078. pmid:25879221
  53. 53. Ross JA, Urton JR, Boland J, Shapiro MD, Peichel CL (2009) Turnover of sex chromosomes in the stickleback fishes (gasterosteidae). PLoS Genet 5: e1000391. pmid:19229325
  54. 54. Miura I (2007) An evolutionary witness: the frog Rana rugosa underwent change of heterogametic sex from XY male to ZW female. Sex Dev 1: 323–331. pmid:18391544
  55. 55. Kirkpatrick M (2010) How and why chromosome inversions evolve. PLoS biology 8: e1000501. pmid:20927412
  56. 56. Rogers RL, Bedford T, Hartl DL (2009) Formation and longevity of chimeric and duplicate genes in Drosophila melanogaster. Genetics 181: 313–322. pmid:19015547
  57. 57. Lynch M (2007) The origins of genome architecture. Sunderland, Mass.: Sinauer Associates. xvi, 494 p. p.
  58. 58. Kennison JA (1981) The genetic and cytological organization of the Y-chromosome of Drosophila melanogaster. Genetics 98: 529–548. pmid:17249098
  59. 59. Gatti M, Pimpinelli S (1983) Cytological and genetic analysis of the Y-chromosome of Drosophila melanogaster .1. Organization of the fertility factors. Chromosoma 88: 349–373.
  60. 60. Allen SL, Delaney EK, Kopp A, Chenoweth SF (2017) Single-molecule sequencing of the Drosophila serrata genome. G3 7: 781–788. pmid:28143951
  61. 61. Goldstein G (2016) Identification of sequences linked to the Y chromosome in the genus Drosophila [MSc thesis]. Rio de Janeiro, Brazil: Federal University of Rio de Janeiro. 70 p.
  62. 62. Reugels AM, Kurek R, Lammermann U, Bunemann H (2000) Mega-introns in the dynein gene DhDhc7(Y) on the heterochromatic Y chromosome give rise to the giant Threads loops in primary spermatocytes of Drosophila hydei. Genetics 154: 759–769. pmid:10655227
  63. 63. Larracuente AM, Clark AG (2014) Recent selection on the Y-to-dot translocation in Drosophila pseudoobscura. Mol Biol Evol 31: 846–856. pmid:24390701
  64. 64. Gao J, Watabe H, Aotsuka T, Pang J, Zhang Y (2007) Molecular phylogeny of the Drosophila obscura species group, with emphasis on the Old World species. BMC Evol Biol 7: 87. pmid:17555574
  65. 65. Alkan C, Sajjadian S, Eichler EE (2011) Limitations of next-generation genome sequence assembly. Nat Methods 8: 61–65. pmid:21102452
  66. 66. van der Linde K, Houle D, Spicer GS, Steppan SJ (2010) A supermatrix-based molecular phylogeny of the family Drosophilidae. Genet Res (Camb) 92: 25–38.
  67. 67. Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): the role of morphology in the postgenomic era. Systematic Entomology 38: 349–364.
  68. 68. Vibranovski MD, Koerich LB, Carvalho AB (2008) Two new Y-linked genes in Drosophila melanogaster. Genetics 179: 2325–2327. pmid:18660539
  69. 69. Rose TM, Henikoff JG, Henikoff S (2003) CODEHOP (COnsensus-DEgenerate hybrid oligonucleotide primer) PCR primer design. Nucleic Acids Research 31: 3763–3766. pmid:12824413
  70. 70. Oliveira DC, Almeida FC, O'Grady PM, Armella MA, DeSalle R, et al. (2012) Monophyly, divergence times, and evolution of host plant use inferred from a revised phylogeny of the Drosophila repleta species group. Mol Phylogenet Evol 64: 533–544. pmid:22634936
  71. 71. Ford MJ, Aquadro CF (1996) Selection on X-linked genes during speciation in the Drosophila athabasca complex. Genetics 144: 689–703. pmid:8889530
  72. 72. Huey RB, Moreteau B, Moreteau JC, Gibert P, Gilchrist GW, et al. (2006) Sexual size dimorphism in a Drosophila clade, the D. obscura group. Zoology (Jena) 109: 318–330.
  73. 73. Sturtevant HA, Novitski E (1941) The homologies of the chromosome elements in the genus Drosophila. Genetics 26: 517–541. pmid:17247021