1471-2164-14-623 1471-2164 Research article <p>Comparative genomics of emerging pathogens in the <it>Candida glabrata</it> clade</p> GabaldónTonitoni.gabaldon@crg.eu MartinTiphainetiphainemartin@sfr.fr Marcet-HoubenMarinammarcet@crg.es DurrensPascalpascal.durrens@labri.fr Bolotin-FukuharaMoniquemonique.bolotin@igmors.u-psud.fr LespinetOlivierolivier.lespinet@igmors.u-psud.fr ArnaiseSylviesylvie.arnaise@igmors.u-psud.fr BoisnardStéphaniestephanie.boisnard@u-psud.fr AguiletaGabrielagabriela.aguileta@crg.eu AtanasovaRalitsarali_bas@yahoo.fr BouchierChristianebouchier@pasteur.fr CoulouxArnaudacouloux@genoscope.cns.fr CrenoSophiescreno@pasteur.fr Almeida CruzJosejoseaccruz@gmail.com DevillersHugohugo.devillers@grignon.inra.fr Enache-AngoulvantAdelaadela.angoulvant@gmail.com GuitardJuliettejulietteguitard@gmail.com JaouenLaurelaure91940@gmail.com MaLaurencelaurence.ma@pasteur.fr MarckChristianchristian.marck@orange.fr NeuvégliseCécilencecile@grignon.inra.fr PelletierEricericp@genoscope.cns.fr PinardAmélieamelie.pinard@gmail.com PoulainJuliepoulain@genoscope.cns.fr RecoquillayJulienj.recoquillay@yahoo.fr WesthofEricE.Westhof@ibmc-cnrs.unistra.fr WinckerPatrickpwincker@genoscope.cns.fr DujonBernardbdujon@pasteur.fr HennequinChristophechristophe.hennequin@laposte.net FairheadCécilececile.fairhead@u-psud.fr

Bioinformatics and Genomics Programme, Centre for Genomic Regulation (CRG) and UPF, Doctor Aiguader, 88, Barcelona, 08003, Spain

Université de Bordeaux 1, LaBRI, INRIA Bordeaux Sud-Ouest (MAGNOME), Talence, F-33405, France

Institut de Génétique et Microbiologie, UMR8621 CNRS-Université Paris Sud, Bât 400, UFR des Sciences, Orsay Cedex, F 91405, France

APHP, Hôpital St Antoine, Service de Parasitologie-Mycologie, and UMR S945, Inserm, Université P. M. Curie, Paris, France

Département Génomes et Génétique, Institut Pasteur, Plate-forme Génomique, rue du Dr. Roux, Paris, F-75015, France

CEA, IG, DSV, Genoscope, 2 rue Gaston Crémieux, Evry Cedex, 91057, France

Architecture et Réactivité de l‘ARN, Institut de Biologie Moléculaire et Cellulaire du CNRS, Université de Strasbourg, Strasbourg Cedex, F-67084, France

Institut de biologie et technologies de Saclay (iBiTec-S), Gif-sur-Yvette cedex, 91191, France

INRA, UMR 1319 Micalis, Thiverval-Grignon, F-78850, France

Institut Pasteur, Unité de Génétique moléculaires des levures, UMR3525 CNRS, UFR927, Université P. M. Curie, 25 rue du Docteur Roux, Paris Cedex15, F75724, France

APHP, Hôpital Bicêtre, Service de Microbiologie, Paris, France

Present adress: Champalimaud Foundation, Av. Brasília, Lisboa, 1400-038, Portugal

Comparative Genomics Group, CRG-Centre for Genomic Regulation, Doctor Aiguader, 88, Barcelona, 08003, Spain

BMC Genomics
<p>Comparative and evolutionary genomics</p>
1471-2164 2013 14 1 623 http://www.biomedcentral.com/1471-2164/14/623 2403489810.1186/1471-2164-14-623
84201331720131492013 2013Gabaldón et al.; licensee BioMed Central Ltd.This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Candida glabrata Fungal pathogens Nakaseomyces Yeast genomes Yeast evolution

Abstract

Background

Candida glabrata follows C. albicans as the second or third most prevalent cause of candidemia worldwide. These two pathogenic yeasts are distantly related, C. glabrata being part of the Nakaseomyces, a group more closely related to Saccharomyces cerevisiae. Although C. glabrata was thought to be the only pathogenic Nakaseomyces, two new pathogens have recently been described within this group: C. nivariensis and C. bracarensis. To gain insight into the genomic changes underlying the emergence of virulence, we sequenced the genomes of these two, and three other non-pathogenic Nakaseomyces, and compared them to other sequenced yeasts.

Results

Our results indicate that the two new pathogens are more closely related to the non-pathogenic N. delphensis than to C. glabrata. We uncover duplications and accelerated evolution that specifically affected genes in the lineage preceding the group containing N. delphensis and the three pathogens, which may provide clues to the higher propensity of this group to infect humans. Finally, the number of Epa-like adhesins is specifically enriched in the pathogens, particularly in C. glabrata.

Conclusions

Remarkably, some features thought to be the result of adaptation of C. glabrata to a pathogenic lifestyle, are present throughout the Nakaseomyces, indicating these are rather ancient adaptations to other environments. Phylogeny suggests that human pathogenesis evolved several times, independently within the clade. The expansion of the EPA gene family in pathogens establishes an evolutionary link between adhesion and virulence phenotypes. Our analyses thus shed light onto the relationships between virulence and the recent genomic changes that occurred within the Nakaseomyces.

Sequence Accession Numbers

Nakaseomyces delphensis: CAPT01000001 to CAPT01000179

Candida bracarensis: CAPU01000001 to CAPU01000251

Candida nivariensis: CAPV01000001 to CAPV01000123

Candida castellii: CAPW01000001 to CAPW01000101

Nakaseomyces bacillisporus: CAPX01000001 to CAPX01000186

Background

Opportunistic fungal pathogens have become a major source of life-threatening nosocomial infections. This situation is partly explained by modern medical progress, relying on large-spectrum antibiotics, immunosuppressive chemotherapy, and devices such as catheters, all of which have been shown to predispose to invasive candidiasis 1 .

Among the emerging fungal pathogens, the incidence of Candida glabrata has progressively increased, and it is currently the second or third most prevalent cause of candidiases. Despite its name, this yeast is phylogenetically closer to the model yeast Saccharomyces cerevisiae than to C. albicans 2 , and is part of the Nakaseomyces genus. This genus originally included three other species of yeasts isolated only from the environment, namely Nakaseomyces (Kluyveromyces) delphensis, Candida castellii and Nakaseomyces (Kluyveromyces) bacillisporus 3 . Recently however, two pathogens have been added to the genus, Candida nivariensis and Candida bracarensis 4 5 . Because routine phenotypic tests, such as biochemical identification methods, are unable to identify these newly described species, leading most often to misidentification as Zygosaccharomyces (CH and AEA, unpublished) their true clinical relevance may be underestimated. Recently, collections of clinical isolates, phenotypically identified as C. glabrata, were screened with molecular methods, and C. bracarensis and C. nivariensis were found to represent less than 2.2% and less than 0.1% of the strains, respectively, with prevalence possibly varying across countries 6 7 . Interestingly, although C. glabrata is considered a commensal of the human gut 8 , the ecological niches of C. bracarensis and C. nivariensis remain unknown. Of note, C. nivariensis has been isolated from flowering plants in Australia 9 , pointing to the possibility that this species may colonize humans from an environmental source.

As it is often the case in fungi, loss, or scarcity of sexual reproduction is associated to species isolated in human patients. Nonetheless, the Nakaseomyces comprise at least one known “environmental” species in which no sex has been observed, C. castellii 10 . MAT-like loci and the HO gene, the key player of mating-type switching in S. cerevisiae 11 , were known to be conserved in C. glabrata and N. delphensis 12 .

The genomic sequence of C. glabrata has been available since 2004 2 , and its comparison to S. cerevisiae has served to discuss possible genomic and metabolic features related to the pathogenic nature of the former 13 . However, it was as yet unclear whether some of these features were also shared by other Nakaseomyces and how their presence actually related to the ability of the different species to become human pathogens. In the case of Candida albicans, this has been explored by sequencing several of its close relatives 14 . To gain a similar insight into the specific features of C. glabrata and their relation to pathogenicity, we now report the complete sequencing of the five other known species in the Nakaseomyces group.

Our results show that all Nakaseomyces nuclear genomes are small, transposon-free and contain significantly less genes than S. cerevisiae. This is in contrast to their mitochondrial genomes which, with the exception of C. glabrata, are large and invaded by palindromic putative mobile elements, the GC inserts 15 . Loss of genes involved in several metabolic pathways as well as loss or amplification of some gene families, are shared by most, sometimes all, Nakaseomyces species, although some remain species-specific. Our molecular phylogenetic analysis supports the phylogeny of the Nakaseomyces, as published by Kurtzman, 3 ie all these species can be grouped together as a new genus with a single common ancestor. We also confirm that the genus can be clearly subdivided into two main lineages, where the lineage containing C. castellii and N. bacillisporus has followed a very divergent evolutionary path. The second group, which we will refer to as the ‘glabrata group’, contains the three pathogenic species and N. delphensis, which is more closely related to the two recently identified pathogens. This depicts a complex scenario suggesting multiple independent events of emergence of pathogenesis within this lineage, and the presence of a genomic repertoire that may facilitate the emergence of pathogenicity towards humans. Altogether our comparative analyses have enabled us to trace, at high levels of resolution, the genomic changes that occurred within this group and discuss how they relate to the pathogenic ability of the different species.

Results

Genome assemblies

Sequencing of the type strains of the five Nakaseomyces species; Candida nivariensis CBS9983, Candida bracarensis CBS10154, Nakaseomyces (Kluyveromyces) delphensis CBS2170, Candida castellii CBS4332, and Nakaseomyces (Kluyveromyces) bacillisporus CBS7720, was performed at Genoscope (Evry, France), using a combination of Illumina and 454 technologies (see Methods). Genome data has been deposited at the EMBL.

Final assemblies showed a close correspondence between scaffolds and chromosomes (Additional file 1), and were annotated for coding and non-coding genes (see Methods). Flow cytometry results show that species are haploid, except N. bacillisporus which is diploid (Additional file 2). All species have an haploid genome size of 10 to 12 Mb. Chromosome numbers, as estimated by Pulsed-Field Gel Electrophoresis (not shown), range from eight in C. castellii, the smallest yet recorded number of chromosomes in post-WGD yeasts 16 , to fifteen in N. bacillisporus; with the ‘glabrata group’ exhibiting the least variation (10–13 chromosomes). Centromeres, similar in structure to S. cerevisiae’s, i.e. composed of three short “centromere defining elements”, CDEI, II and III 17 ; were identified in all species but not in all scaffolds. Telomeric repeats identical to those in C. glabrata and the putative telomerase RNA-component were found in the ‘glabrata group’.

<p>Additional file 1</p>

Assembly data.

Click here for file

<p>Additional file 2</p>

Flow cytometry of the Nakaseomyces. Species names are indicated above each panel. 1C, 2C, 4C indicate peaks corresponding to the DNA content of, respectively, one, two and four haploid genomes.

Click here for file

Several ncRNA genes are known to be surprisingly large in C. glabrata, such as the RPR1 gene of RNAse P 18 and the above-mentioned TLC1 gene 19 . This tendency to exceptionally large ncRNAs seems to be general in the Nakaseomyces, such as the 1368 nt-long RPR1 gene in N. delphensis (only 369 nt-long in S. cerevisiae), and the 1937 nt-long U1 snRNA gene in C. castellii (568 nt-long in S. cerevisiae). Other structural genomic features are discussed in the supplementary results (see Additional file 3). None of the species contain detectable active transposons in their nuclear genome.

<p>Additional file 3</p>

Supplementary text containing additional results and references. Table S1. tRNA genes in the Nakaseomyces genomes. Table S2. GC inserts in the mitochondrial genomes of C. bracarensis and C. nivariensis.

Click here for file

Phylogeny of Nakaseomyces

In order to clarify the phylogenetic relationships of the newly-sequenced Nakaseomyces and 17 other Saccharomycotina, we used two alternative phylogenomics approaches, namely i) a Maximum Likelihood (ML) analysis of a concatenated alignment of 603 protein families that have one-to-one orthologs in all the species considered, and ii) a super-tree approach based on the analysis of 4,965 individual gene trees, which finds the species topology that is most parsimonious in terms of implied duplication events in all the individual gene trees 20 . Both approaches yielded the same, highly resolved topology (Figure 1), which is largely congruent, for the shared species, with our current understanding of Saccharomycotina phylogeny 21 .

<p>Figure 1</p>

Maximum likelihood species tree of 22 Saccharomycotina species.

Maximum likelihood species tree of 22 Saccharomycotina species. The tree was reconstructed based on the analysis of a concatenated alignment of one-to-one orthologs of 603 widespread genes. Species names in red and with an asterisk indicate human fungal pathogens. Underlined species names correspond to the newly sequenced Nakaseomyces species. Important evolutionary events such as the Whole Genome Duplication (WGD) or the genetic code transition in the Candida clade (CTG) are marked on the tree. All aLRT-based supports were maximal and a single node with a bootstrap support below 100% is indicated. This topology is also the most parsimonious in terms of inferred duplications in 4,965 individual gene phylogenies, as assessed by a Gene Tree Parsimony approach implemented in duptree 20.

This topology supports the the existence of the Nakaseomyces genus, and defines two clear sub-groups separated by an ancient split. The first group comprises the highly divergent C. castellii and N. bacillisporus, whereas the second group (i.e. the ‘glabrata group’) includes the three pathogenic species and N. delphensis. The average protein identity between orthologs of species in the different sub-groups ranges from 51 to 53% (Additional file 4), which is similar to that of orthologs in the C. castellii/N. bacillisporus (53%) and C.glabrata/S.cerevisiae (54%) pairs, pointing to large levels of divergence. The ‘glabrata group’, is more compact and shows higher identity levels (77-88%). Notably, the two newly identified pathogens, C. nivariensis and C. bracarensis, are both closer to N. delphensis than to the most frequently isolated pathogen, C. glabrata (Figure 1). Thus the emerging picture for the appearance of pathogenesis is complex, with plausible alternative scenarios involving either gain of the ability to infect humans at the base of the sub-clade followed by loss of the trait in N. delphensis or three independent events of acquisition of pathogenicity. These possibilities will be further discussed below.

<p>Additional file 4</p>

Pair-wise species identity between orthologous protein-coding genes. Each histogram represents the numbers of orthologous gene pairs according to their percentage of identity.

Click here for file

The degree of synteny, i.e. gene order conservation between species, was found to correlate with phylogenetic proximity, with the highest conservation occurring between C. bracarensis, C. nivariensis and N. delphensis (Additional file 5). Conservation of synteny of the Nakaseomyces in general, relative to S. cerevisiae is low. Detailed analysis of syntenic regions will be presented elsewhere (HD, TG, CF, in preparation).

<p>Additional file 5</p>

Number of synteny blocks according to the mean length of synteny blocks. All pairwise comparisons of the genomes of the Nakaseomyces and S. cerevisiae (out-group) are shown. Genomes were divided into three groups according to their location in the phylogenetic tree (Figure 1): the ‘glabrata group’ (red); C. castellii and N. bacillisporus (green); S. cerevisiae (blue). Each dot corresponds to one pairwise comparison and is colored according to the groups of the two compared genomes.

Click here for file

Mating types

As is the case for many fungal species described as asexual, genes involved in sexual reproduction are known to be conserved in C. glabrata 22 23 24 . In particular, although mating has never been reported in C. glabrata, the two additional MAT cassettes, HMRa and HMLalpha, and the HO gene are present in its genome, and haploid isolates of both mating types are found. In N. delphensis, also a mainly haploid species, these elements are also conserved 12 23 and mating-type switch may occur in culture 12 . For the remaining four species, three are considered asexual and mainly haploid, but the last one, N. bacillisporus, is considered to be diploid and homothallic 25 . As mentioned above, ploidy was confirmed by flow cytometry. All species contain a well-conserved homolog of the HO gene (Additional file 6), the most diverged encoded protein being the one from N. bacillisporus, which exhibits a C-terminal extension rich in proline and serine.

<p>Additional file 6</p>

Alignment of putative Ho proteins. “LAGLIDADG” motifs and Nuclear Localization Signals are boxed.

Click here for file

In all species, additional sequencing was needed in order to obtain the cassettes. In N. delphensis, N. bacillisporus and C. castellii, amplification of the MAT-like cassette yielded both a and alpha fragments. This raises the intriguing possibility that C. castellii switches mating types in culture, whilst having no described sexual cycle, or that it is sexual and goes through a diploid phase. This will need further experimental analysis. Figure 2 shows which cassettes are currently identified in these genomes. Genes within cassettes, a1, alpha1 and alpha2 are identified in all species, as is the Ho site, which can be recognised at the YZ junction. In cases where the three cassettes are identified, their configuration is apparently similar to C. glabrata’s: both HML- and MAT-like cassettes are on the same scaffold, while the HMR-like cassette is on a different one, except for N. delphensis, where the three cassettes are on the same scaffold. Thus, Nakaseomyces species follow the rule of conservation of HML and MAT on the same chromosome, noticed by Gordon et al. 26 .

<p>Figure 2</p>

MAT-like cassettes in the sequenced genomes.

MAT-like cassettes in the sequenced genomes.HMR-, HML- and MAT-like are represented as rectangles, with red ones containing a-type information and blue ones containing alpha-type information. MAT cassettes of each type are shown for N. bacillisporus, which is diploid, and bicolored MAT cassettes in haploid species indicate possible switching in culture. Approximate coordinates on corresponding scaffold are shown below each cassette.

C. nivariensis has two HMR-like cassettes, i.e. cassettes that contain type a information, and that are present in addition to the MAT locus, a situation already noted in other species 26 . Experimental testing of these cassettes will reveal what role, if any, these extra loci have, in organisms where sexual reproduction has not been characterized yet.

Variations in gene repertoires

The total numbers of protein-coding genes range from 5400 to 5900, which is similar to what is found in C. glabrata 2 , and lower than in S. cerevisiae. Indeed, the number of true protein-coding genes in S. cerevisiae is estimated at around 5800, but this rises to around 6600 when dubious and other non-experimentally characterized ORFs are included, a figure more comparable to our predicted gene sets (SGD, 12 January 2012, http://www.yeastgenome.org, and Additional file 7). To assess the coverage of the predicted gene repertoires we tested for the presence of a set of 2,007 protein families previously found to be widespread in Saccharomycotina 21 . Our proteomes contained 99.4-100% of this core dataset, attesting for a high coverage in our procedures.

<p>Additional file 7</p>

Histograms of Nakaseomyces genes according to the length of encoded proteins.

Click here for file

Intron-containing protein-coding genes are far fewer in C. glabrata than in S. cerevisiae (129 vs 287, 27 ). This paucity of introns is shared by all Nakaseomyces, which have intron counts lower than 230. Although experimental validation is needed, our data point to a remarkable loss of introns in the Nakaseomyces.

To accurately trace the evolution of the genetic repertoires across the Nakaseomyces, and confidently establish orthology and paralogy relationships, we performed an exhaustive phylogenomic analysis that included the reconstruction of ML phylogenies for every gene encoded in the Nakaseomyces genomes (i.e. the phylome). These were used to detect orthology and paralogy relationships 28 , and to map lineage-specific gene duplication 29 and gene loss events onto the species tree. Figures 3 and 4 and Additional file 8 summarize the main findings regarding the presence or absence of genes relevant to central processes. Lower gene numbers in the Nakaseomyces when compared to S. cerevisiae are reflected in the number of gene loss events indicated on the tree from Figure 3.

<p>Figure 3</p>

Summary of the main findings.

Summary of the main findings. The phylogenetic tree represents the evolution of the Nakaseomyces species, using S. cerevisiae as outgroup. Blue numbers indicate the number of genes that are predicted to have been gained at each lineage during the evolution of the Nakaseomyces. Red numbers indicate the yeast genes that have been lost. Species names coloured in red indicate the human pathogens, lighter colouring indicates recently-reported emerging pathogens. Coloured branches can be matched to the corresponding coloured boxes below, which list important events occurring at that lineage in the evolutionary history of Nakaseomyces. (Abbreviations: Ploidy: H: Haplobiontic, D: Diplobiontic; Sexual reproduction: H: homothallic, N.O.: not observed).

<p>Figure 4</p>

Phylogenetic profiles of specific gene families and pathways.

Phylogenetic profiles of specific gene families and pathways. The phylogenetic tree represents the evolution of the Nakaseomyces species, with the pathogens colored. White boxes indicate absence of a particular family or pathway in a given species, while numbers in colored boxes indicate the number of paraloguous copies of that gene family or the number of components of a given pathway. Intensity of the colors is proportional to the number of paralogs present.

<p>Additional file 8</p>

Genes from S. cerevisiae that are absent from C. glabrata and their absence/presence in the Nakaseomyces.

Click here for file

In particular, we have paid special attention to the retained copies of so-called ohnologs (i.e. gene duplicate pairs originated during the WGD event; 30 ). In S. cerevisiae, only 551 ohnologous pairs are left in the contemporary genome, as most such pairs have lost one member 31 . This is also the case in the Nakaseomyces genomes, as in all post-WGD species. The exact complement of ohnologs retained in duplicates varies across species and is likely to reflect particular physiological differences.

A noteworthy example comes from central carbon metabolism: it has been suggested that the six ohnologous gene pairs encoding glycolytic enzymes, such as ENO1/ENO2, and PYC1/PYC2, conserved in S. cerevisiae and C. glabrata, play an important role in the Crabtree effect, ie fermentation even in the presence of high glucose and oxygen 32 , a trait shared by all Nakaseomyces 33 34 . The genes involved are also in pairs in the ‘glabrata group’, but, in both C. castellii and N. bacillisporus, the situation is rather different, with single pyruvate kinase, enolase, and Glyceraldehyde-3-phosphate dehydrogenase genes (Figure 4, and Additional file 9). Furthermore, C. castellii has two hexokinase genes (HXK), as compared to four (two pairs of ohnologs) in the other species. Other features of central carbon metabolism in the Nakaseomyces are mentioned below and in Additional file 3.

<p>Additional file 9</p>

Glycolytic enzymes in the Nakaseomyces.

Click here for file

Comparison of the proteins encoded in C. glabrata and S. cerevisiae’s genomes had revealed several features specific to the former. In particular, since there are fewer genes in C. glabrata than in S. cerevisiae, there was a special interest in specific gene loss. Indeed, there are at least four entire multigenic families which are absent in all Nakaseomyces or represented by a sole member in C. castellii and/or N. bacillisporus: the PHO, SNZ, SNO and PAU families. It is noteworthy that the paralogous PHO family of acid phosphatases has been lost, while the rest of the PHO pathway is conserved. Functional analysis in C. glabrata has shown that the PHO2 gene is present but not essential for regulation, and that the PHO4 gene is poorly conserved at the sequence level, but is functional 35 . Phosphate-starvation induced phosphatase activity in C. glabrata has been identified and is encoded by PMU1 homologs in tandem ( 36 , and see section on tandem arrays). The SNZ and SNO gene families are poorly characterized in S. cerevisiae, but their expression is known to be induced in stationary phase. Finally, the PAU family consists of at least 20 subtelomeric genes in S. cerevisiae, possibly involved in anaerobiosis 37 and similar to two other gene families, DAN and TIR, that encode cell-wall mannoproteins. There are no homologs to any of these genes in C. castellii and N. bacillisporus, while C. glabrata and C. nivariensis contain two copies of DAN-like genes, and the remaining species seem to contain only one homologous copy.

Also of particular interest was the loss of genes involved in de novo biosynthesis of nicotinic acid (BNA), which was hypothesized to result from the adaptation of C. glabrata to its human host 38 . Comparison to the newly sequenced Nakaseomyces has shown that the lack of BNA genes is common to them all (Additional file 8), regardless of their habitat, suggesting that the loss of this pathway is an ancient event. Notably, loss of BNA genes has also been observed in other, more distant species, such as Kluyveromyces lactis and other species, and seems to be a volatile trait 39 . In fact, all gene losses observed in C. glabrata relative to S. cerevisiae are shared by the ‘glabrata group’, whereas in the two other species gene absence is variable. A remarkable example concerns the genes necessary for allantoin catabolism. In S. cerevisiae, the subtelomeric DAL (

    d
egradation of
    al
lantoin) cluster consists of six genes, and was formed in the ancestor to S. cerevisiae and Naumovia castellii, but lost in C. glabrata 40 . The DAL cluster is absent from the ‘glabrata group’, but is present in both C. castellii and N. bacillisporus. Intriguingly, in these genomes, the cluster contains another gene involved in allantoin catabolism, DAL5, which, in S. cerevisiae, is not linked to the cluster (Additional file 10).

<p>Additional file 10</p>

The DAL cluster. The cluster from S. cerevisiae is shown at top. The cluster containing the additional DAL5 gene in C. castellii and N. bacillisporus is shown below, using the gene nomenclature from S. cerevisiae. In these two genomes, the cluster differs only by the synteny on the left. Genes are represented by arrows, genes in black are DAL genes.

Click here for file

Other notable examples of gene gain/loss involve the translation machinery: both the number of ohnologous ribosomal protein (RP) gene pairs and the RP gene regulators, CRF1 and SFP1, vary between species, with no correlation between absence/presence of regulator genes and number of ohnologous RP gene pairs (Figure 4, Additional files 3 and 11).

<p>Additional file 11</p>

Duplicated ribosomal protein genes in the Nakaseomyces.

Click here for file

Central carbon metabolism again provides examples of gene gain/loss events; all Nakaseomyces genomes contain only two copies of ADH genes, which, both by similarity and by conservation of synteny, correspond to orthologs of ADH1 and ADH3, the S. cerevisiae genes that encode, respectively, the cytoplasmic and the mitochondrial activities converting acetaldehyde into ethanol. ADH2, which in S. cerevisiae is specialized in the conversion of ethanol to acetaldehyde, has no ortholog in the Nakaseomyces. It is possible that bi-directional activities exist, or that alternative enzymes take over this conversion (i.e. co-option). For example, S. cerevisiae has additional alcohol dehydrogenase genes, in particular the family of aryl alcohol dehydrogenases encoded by seven subtelomeric AAD genes and the non-subtelomeric YPL088W gene 41 . C. glabrata possesses an array of three such genes in tandem, while other Nakaseomyces have several dispersed copies, except C. castellii which harbors a single such gene. Experimental analysis is needed to tell which enzymes catalyze which reactions in Nakaseomyces, and even in S. cerevisiae, in which enzymatic activities are still in the process of being characterized 42 . As for anaerobic growth, all Nakaseomyces have the ability to grow in micro-aerobiosis, as tested by standard laboratory methods (not shown). Two pairs of regulators are described as essential to anaerobiosis in S. cerevisiae; the ECM22/UPC2 pair and the SUT1/SUT2 pair. These genes are conserved in all Nakaseomyces except C. castellii. There are also pairs of genes that differ by their expression under aerobic and anaerobic growth, such as COX5A/COX5B and of CYC1/CYC7 in S. cerevisiae. In contrast, all Nakaseomyces retain a single member of each of these ohnologous pairs, raising the question of their regulation.

Genes involved in virulence

In C. glabrata, the EPA genes, a family of glycosyl-phosphatydylinositol (GPI)-anchored cell-wall protein (CWP) genes 43 , are the best characterized genes involved in adhesion to human epithelia 44 45 , an ability associated to virulence in diverse pathogens 46 . Notably, our search for homologs of C. glabrata EPA genes in the newly sequenced Nakaseomyces (see methods), revealed higher numbers of such genes in the three pathogenic species. More specifically, in the most prevalent pathogen, C. glabrata, the type strain harbors 18 members of this family. Seven new variants of EPA genes are found in a different C. glabrata strain, BG2 (data kindly provided by Brendan Cormack). The other two pathogenic Nakaseomyces, C. bracarensis and C. nivariensis have respectively, 12 and 9 members of the EPA family. In contrast, the non-pathogenic N. delphensis harbors a single copy. These differences cannot be attributed to differences in the quality of assemblies, which were similar in all newly sequenced species. Of the two remaining Nakaseomyces species, only C. castellii contains three homologs of the EPA genes, while N. bacillisporus presented only one distant homolog that clustered with PWP (PA-14 containing Wall Protein, adhesin gene) and adhesin-like protein genes in C. glabrata. These proteins, more similar to the floculin homologs in S. cerevisiae, are only distantly related to Epa proteins 47 48 . A closer inspection of the corresponding trees in the phylome and of a composite tree constructed with all members of the identified EPA-like members (Figure 5a), revealed that a significant fraction of the adhesin gene copies in C. glabrata emerged from lineage-specific duplications in this species, whereas many duplications in C. bracarensis and C. nivariensis are shared. This independent expansion of EPA-like genes in C. glabrata and the emerging pathogens supports the idea of an independent emergence of pathogenesis and may explain the important differences in prevalence across these species. Strikingly, the non-pathogenic member of the ‘glabrata group’, N. delphensis possesses a sole representative of this family, although the phylogenetic scenario implies that the common ancestor of this species and the two emergent pathogens would have had a higher number of copies. This scenario implies specific losses of EPA-like genes in N. delphensis, and a possible secondary loss of adhesion capabilities in this species. Interestingly, our search for adhesin-like proteins uncovered the presence of a group of four proteins in N. delphensis with poor similarity with the EPA-like family (Figure 5c). Whether this group represents highly divergent members -or even pseudogenes- of the EPA-like family in N. delphensis, remains to be further investigated. In any case, their high levels of sequence divergence, particularly at the N-terminal region known to harbor the ligand-binding domain in Epa 49 , suggest they cannot be functionally equivalent.

<p>Figure 5</p>

Trees of the three groups of adhesins containing more than one member.

Trees of the three groups of adhesins containing more than one member. Phylogenetic trees were reconstructed using the same approach as for the trees present in the phylome, and are visualized using ETE 50, A) proteins grouped with the known C. glabrata EPA genes, B) proteins similar to the yeast Flo protein, C) group of unknown adhesins, and D) sequences related to AWP adhesins. Labels indicate protein names and are colored depending on the species they belong to: C. glabrata in dark blue, a second strain of C. glabrata (BG2) in light blue, C. bracarensis in red, C. nivariensis in green, N. delphensis in purple, N. bacillisporus in yellow and C. castellii in orange. Logos at the lower right part of the image represent the conservation in each of the adhesin groups of sites that are considered important in the structure of EPA151.

Other genes shown to be involved in virulence in C. albicans and/or C.glabrata, such as the phospholipase B gene, or the Super Oxide Dismutase genes (SOD genes), exhibit variable presence/absence in the Nakaseomyces, with no obvious correlation to pathogenicity of the species.

Genes in tandem arrays

Tandem arrays are a special genomic arrangement of certain gene families that imply specific amplification mechanisms and positive selection. They can either be shared by all members of a species, such as the globin genes in mammals 52 , or display polymorphism in terms of number of copies, such as the CUP1 gene in S.cerevisiae, whose amplification is positively selected in copper-containing medium 53 . A few dozen tandems were identified in the Nakaseomyces genomes (HD, TG and CF, ms in preparation). The species with the largest tandem arrays is C. glabrata, with two arrays of eight genes each (Figure 6), the MNT3 array (shown to be variable in different strains, 54 ) and the YPS array 55 , and an array of five genes, coding for a protein of unknown function predicted to be involved in carbohydrate metabolism. None correspond to tandem arrays of more than two genes in other Nakaseomyces. Two other arrays of three genes are specific to C. glabrata, the aryl alcohol dehydrogenase gene array, and the PMU1 array. These findings indicate that most tandem arrays found in C. glabrata originated specifically in this lineage. This, together with the finding that some of these arrays are variable across strains, suggests that the amplification of these families may be driven by (directional) selection. Functions encoded by these families, are thus good candidates for finding possible physiological advantages underlying the success of C. glabrata as an opportunistic pathogen. In the case of the YPS cluster, the genes have been shown to be involved in virulence 55 . PMU1 encodes the phosphate-starvation inducible phosphatase activity, which has been hypothesized to be a specific adaptation of C. glabrata to its mammalian host 36 .

<p>Figure 6</p>

Tandem repeats of eight genes identified in C. glabrata and their homologs in the ‘glabrata group’.

Tandem repeats of eight genes identified in C. glabrata and their homologs in the ‘glabrata group’. A: MNT3 homologs in C. glabrata are present in a tandem repeat of eight genes and in a tandem of two genes, with additional isolated copies. Closest species exhibit a tandem repeat of two genes with mixed synteny at the borders. B: YPS homologs in C. glabrata are present in a tandem repeat of eight genes, with additional isolated copies. Closest species exhibit a tandem repeat of two genes with conserved synteny at the borders.

Analysis of evolutionary rates in the lineages leading to the ‘glabrata group’

As mentioned above, the ‘glabrata group’ contains several species with the ability to infect humans. We have previously discussed changes that occurred within this lineage in terms of acquisition and loss of genes. We next set out to investigate whether we could identify signatures of possible positive selection in the form of genes with accelerated rates in the branch preceding the diversification of the ‘glabrata group’. For this we focused on 2,153 genes predicted as one-to-one orthologs shared by all Nakaseomyces species and S. cerevisiae. Using the species’ phylogeny, we used a likelihood ratio test (LRT) to compare two nested rate models (Additional file 12). For 991 of the 2153 genes (43%), a model including a different rate in the branch leading to the ‘glabrata group’ was favored, and in 35 of them there was evidence for accelerated rates, suggestive of positive selection (i.e. d N /d S rate >1, with an average of 3). In contrast, we did not find a significant rate acceleration in either the S. cerevisiae, N. bacillisporus or C. castelli lineages, nor within the ‘glabrata group’ itself, where the average d N /d S rate was 0.04 indicating high levels of purifying selection. These findings suggest that prior to the diversification of the ‘glabrata group’, there was an increase in the non-synonymous substitution rate in at least 35 genes (1.6% of the tested genes). Among these we did not find families that have been related to pathogenesis in C. albicans. Finally, we focused on the ‘glabrata group’ to identify genes accelerated in parallel in the three lineages leading to the three pathogenic species or exclusively in the C. glabrata lineage. We identified 94 proteins with evidence for different selective constrains in pathogenic species and N. delphensis, although this difference was mostly due to stronger purifying selection, rather than positive selection, in the pathogens. With respect to the C. glabrata specific branch, only four genes presented d N /d S >1. Overall, these results would suggest a burst of nonsynonymous substitution rates in a significant number of genes preceding the divergence of the ‘glabrata group’ and a subsequent stasis within the group itself. Nevertheless, genes that are positively selected in these lineages constitute good candidates for testing potential roles in virulence.

<p>Additional file 12</p>

Scenarios for positive selection tests.

Click here for file

Discussion

Even though experimental data will be needed to clarify the relationship between genome content and adaptation to the environment, we show that many gene variations observed occur in families of genes encoding cell wall proteins (PAU, EPA, and other ministallite-containing genes), as well as proteins involved in carbohydrate metabolism. These classes of genes have already been shown to be involved in adaptation of yeast species to particular biotechnological niches such as the MEL and MAL genes in Saccharomyces species, 56 57 , the SUL genes in lager yeasts 58 or in adaptation to the human host such as the EPA genes in C. glabrata 55 and the ALS genes in C. albicans 59 . Some of these genes fit the definition of contingency genes, i.e. genes that encode products that mediate the cell’s response to its environment and that evolve faster than allows the average mutation rate of other genes 60 . Rapid evolution is supposed to be facilitated by internal repeats, such as within the EPA genes, and by subtelomeric location, situations which enhance recombination frequency and generate new alleles and possibly new genes.

Our data firmly support the existence of the Nakaseomyces genus, but nonetheless, two of the species, C. castellii and N. bacillisporus, have diverged in many ways, first by the low conservation of their orthologous genes, when compared to the four other Nakaseomyces, and also by their paucity of tandems, and by their frequent variation in gene numbers compared to the ‘glabrata group’ (such as in “ohnologous” ribosomal protein gene pairs). Furthermore, our analyses shows that the two emerging pathogens, C. nivariensis and C. bracarensis, are most closely related to the non-pathogenic N. delphensis, and all descend from an ancestor that has also given rise to C. glabrata. In accordance to their phylogenetic relationship, most genomic particularities observed in C. glabrata as compared to S. cerevisiae, are also shared by the two emerging pathogens and N. delphensis. Some others, such as the absence of the nicotinic acid synthesis pathway are even common to all Nakaseomyces, and thus represent ancient Nakaseomyces traits, whose origin must have pre-dated the adaptation to the human host of some of the members of the group. Hence, the scenario for the emergence of pathogenicity within the Nakaseomyces, and the underlying genomic and metabolic changes must be reinterpreted in the light of these new genomic data. These findings also highlight the utility of increased taxonomic samplings when correlating genomic and phenotypic differences. Considering that most genomic and metabolic features previously thought to be particular of C. glabrata are shared by the three most closely related Nakaseomyces, of which two are not natural inhabitants of the human gut, the most plausible explanation is that they result from adaptations to environments other than the human gut. The alternative scenario in which human commensalism is an ancestral trait that was lost in C. nivariensis and N. delphensis seems unlikely given the large evolutionary distances involved and the relatively recent origin of our species. Furthermore, since these four related species include one that has never been identified as a human pathogen, and two that have only been recently reported as opportunistic pathogens, the link to the emergence of pathogenesis for their specific genomic features should be indirect. Nevertheless, the presence of three related species able to infect humans within this sub-group of Nakaseomyces is in stark contrast to the almost complete absence of this ability in the rest of the genus as well as in the related Saccharomyces and Kluyveromyces groups. Thus, it seems that some of the particularities shared by C. glabrata and the three related species may represent pre-adaptations (exaptations in evolutionary terms) that may have facilitated, but not directly triggered, the emergence of pathogenecity towards humans. Although further research is needed to identify which traits may have been particularly important for the emergence of pathogenesis within the Nakaseomyces, several of the traits shared by C. glabrata and the three closest relatives are firm candidates. These include genes that likely underwent positive selection specifically in this lineage, of which some have homologs implicated in pathogenesis in C. albicans. In addition, the multiple parallel expansions of the EPA genes, known to be important for the ability of C. glabrata to adhere to human cells, represent a clear example of a genomic change that correlates with the pathogenic trait. Our data supports the fact that the emergence of pathogenesis relies on a combination of genomic alterations, rather than changes in a single gene family.

Undoubtedly, C. glabrata is the most prevalent pathogen among the Nakaseomyces. This increased ability to infect immunocompromised humans is probably related to some of the specific changes observed in the C. glabrata lineage, as compared to the other Nakaseomyces. These include specific losses, expansions of some gene families, particularly as tandem arrays, and even the acquisition of horizontally transferred genes 61 . Remarkably, some of these traits specific to C. glabrata have been related to virulence, such as the largest expansion of gene families involved in cell adhesion (EPA) and in phosphate starvation (PMU1). That the EPA genes have been expanded in C. glabrata independently of the emergent pathogens is consistent with the important differences of prevalence and suggest an independent emergence of increased adherence to the human epithelium, an important virulent trait.

Considering all these data, one can speculate on the possible scenarios depicting the origin of pathogenesis within the Nakaseomyces. To start with, our phylogeny does not support the monophyly of C. glabrata and the two emerging pathogens. Two competing scenarios may explain this pattern: a single origin of pathogenesis towards humans followed by loss of the trait in N. delphensis or, alternatively, at least two independent events of emergence of pathogenesis. Although the first scenario is more parsimonious in terms of the number of phenotypic shifts, the second hypothesis seems more likely in the light of the parallel expansions of the EPA genes, and the fact that the pathogens other than C. glabrata have only been recently identified, suggesting these are recently emerged rather than derived pathogens. A plausible scenario for the emergence of pathogenesis within the Nakaseomyces, compatible with our data, comprises the following steps: an ancestral environmental yeast with specific genomic features, gives rise to species adapted to being commensals of humans, of which some can evolve into opportunistic pathogens. A certain level of adaptation to the mammalian gut may have represented a selective advantage to yeast species that are naturally present in edible parts of plants (i.e. fruits), since this may have facilitated dispersion by the animals consuming the plants. Increased levels of adaptation to the mammalian gut environment may have resulted in species that persist in the gut and gradually adapt to a particular host (e.g. human). Once reached this point, particular features of some species may provide them with the ability to colonize an immuno-compromised host. Such features could be related to the ability to adhere to the host, persist in tissues other than the gut and to overcome the (debilitated) host immune system. Intriguingly, C. nivariensis seems not to be a inhabitant of the human gut, and may colonize human patients from an environmental source. Although further research is needed to clarify this, emergence of pathogenesis from environmental species would suggest that prior commensalism with humans is not a pre-requisite for developing infection capacity in Candida spp. Nevertheless, such species found in the environment may also be associated to other mammals. Clearly, additional data on ecological distribution of these Nakaseomyces species is needed to sort out these alternative hypotheses.

Conclusions

Figure 3 shows a summary of the main findings described in this work. Comparative genomics analyses support the hypothesis that pathogenicity arose several times in the Nakaseomyces, and that this group represents a true genus, with common ancestral traits that may be favorable to adaptation to the human host.

Methods

Strains and DNA preparation

All strains are the type strains of the corresponding species 4 5 8 10 62 63 , obtained from the CBS collection. All media used were prepared as for S. cerevisiae: glucose and glycerol-based complete media, broth and solid. Cultures were performed at 30°C or 37°C, with agitation for broth cultures.

Anaerobic growth was tested by inoculating Sabouraud plates and using the Oxoid™ Anaerogen system 64 . Plates were examined after 120 h of incubation.

Petite mutants were obtained by exposing cells to Ethidium Bromide 65 . Petite mutants completely lacking mitochondrial (mt) DNA (ρ0 mutants) were used for the flow cytometry experiments so that mt DNA did not interfere with measures, since mt DNA content can be quite high in these species.

For the same reason, DNA for sequencing was prepared by standard zymolyase extraction followed by separation on a CsCl gradient with bis-benzimide 66 . The upper mt DNA band was discarded and the lower band used for sequencing. This procedure allowed the mt DNA sequence to be acquired nonetheless at an acceptable fold-coverage.

Ploidy determination by flow cytometry

An aliquot of 4 mL from a fresh yeast culture at 1–2 106 cells/mL is mixed with 9.2 mL of pure ethanol and incubated at 4C overnight. After centrifugation, cells are washed in 50 mM, pH7 Sodium Citrate and resuspended at 108 cells/mL. An aliquot of 200 μL is treated with RNAse by adding 2 μL of a 100 mg/mL solution and incubating 2 hrs at 37°C. Half is then labeled by adding 400 μL of Propidium Iodide at 50 μg/mL in 50 mM Sodium Citrate, and incubating 20mn in the dark. The sample is then ready for the flow cytometer.

Sequence and annotation of protein-coding genes

Sequencing was done by the Genoscope (Evry, France). Briefly, a whole genome shotgun (WGS) strategy associating different types of sequencing technologies was performed for each strain. An mean 24.8 genome equivalent (from 15.4 up to 41) was achieved using 454 GSFlx approach using a mixture of 8 kb mate pairs sequencing and single reads. Genome assembly was performed using 454 Newbler software. Subsequently, a mean 70 genome equivalent coverage for each strain was obtained using Illumina GAIIx technology 36 or 76 bp single reads, and these data were used to correct assembly errors 67

Probably because of the complete sequence identity between segments of the three MAT-like cassettes, most of these loci were absent from the automated assembly generated from the Illumina and 454 reads. Only three loci were present in the assemblies: one HMRa from C. bracarensis and two HMRa from C. nivariensis. We therefore searched for sequence gaps in regions of synteny with the cassettes from C. glabrata. In all cases except one, there was indeed a gap in the region where the cassette was expected to be, by synteny conservation. The HMR cassette from N. bacillisporus is still missing from the assembly. Fragments were amplified by PCR and sequenced by Sanger sequencing.

Annotation of protein-coding genes was performed with an in-house procedure, which consists of two phases: syntactical annotation (prediction and location of protein coding genes), followed by functional annotation of each element based on comparison with known sequences. The first phase calls upon 6 gene prediction algorithms: CONRAD 68 , AUGUSTUS 69 , GETORF 70 , SNAP 71 , GENEMARK 72 , GENEID 73 , using the same training set of gene sequences, which contains genes with and without introns, for those which needed a training step; the intron-containing genes being defined by comparison to intron-bearing genes of C. glabrata and S. cerevisiae. All predictions as well as tBLASTn alignments to proteomes of reference species and Uniprot, and PSI-tBLASTn alignments to PSSM representative of Génolevures protein families are integrated using JIGSAW 74 . Then, all gene models from the 6 prediction algorithms plus JIGSAW are put together and filtered to eliminate gene models having unrealistic introns. The overlap conflicts between elements and validation of gene models are solved by taking into account predicted gene models, other chromosomal elements already validated, and similarity regions, strands and frames. The resulting gene models are then submitted to functional annotation, based on a decision tree inspired by previous semi-automated annotation projects held by the Génolevures Consortium (which used BLASTp alignments to proteomes of reference species and Uniprot).

Identification of centromeres was done by searching with fuzznuc from the Emboss suite 70 , for sequences homologous to S. cerevisiae consensus centromere sequence, ie “[AG]TCA[TC][AG]TG[AC][TC]N(73,167)G[GT]N(7,15)TTCCGAA” 75 , and by manually checking for conservation of synteny in case of multiple hits within a single scaffold. Telomeric repeats were searched for by using the repeat motif from C. glabrata, CAGCACCCAGACCCCA, as blastn queries against the genomic sequences. This also revealed the putative template inside the TLC1 gene.

ncRNA discovery and annotation

tRNA genes were identified by both cloverleaf structure detection 76 and tRNAscan-SE 77 . BLAST 78 and Infernal 79 searches were performed on each of the five Nakaseomyces genomes for other ncRNA genes; using annotated ncRNAs from the Genolevures database 80 as queries for BLAST, and the covariance models from RFam database 81 for the Infernal search. The hits of both searches were combined according to the respective ncRNA family and extended in order to detect contiguous hits corresponding to a same candidate. All hits from the same families were aligned and manually checked. Hits were accepted as candidates if: i) the sequence agrees with known structural features, guiding sequences (for snoRNAs) and conserved sequence motifs for homologous molecules and ii) known synteny was verified.

Phylogenomics

A phylome, the complete collection of phylogenetic trees for each gene in a genome, was reconstructed for each one of the six Nakaseomyces species (five newly sequenced and the reference C. glabrata genome). The phylomes include 16 other species: Saccharomyces cerevisiae, S. mikatae, S. (Naumovia) castellii, S. kluyveri, S. bayanus, Vanderwaltozyma polyspora, Lachancea thermotolerans, Ashbya gossypii, Candida dubliniensis, Kluyveromyces lactis, Candida albicans, Debaryomyces hansenii, Zygosaccharomyces rouxii, Yarrowia lipolytica, Pichia stipitis, Lachancea waltii. Phylomes were reconstructed using the pipeline described in 82 . In brief, for all genes in each Nakaseomyces genome, a Smith-Waterman search 83 was used to retrieve homologs using an e-value cut-off of <10-5, and considering only sequences that aligned with a continuous region representing more than 50% of the query sequence.

Once the sets of homologous sequences were defined, phylogenetic trees were reconstructed as follows. Selected sequences were aligned using three different programs: MUSCLE v3.7 84 , MAFFT v6.712b 85 , and DIALIGN-TX 86 . Alignments were performed in forward and reverse direction (i.e. using the Head or Tail approach 87 ), and the six resulting alignments were combined using M-COFFEE 88 . The resulting combined alignment was subsequently trimmed with trimAl v1.3 89 using a consistency score cutoff of 0.1667 and a gap score cutoff of 0.9.

The selection of the evolutionary model best fitting each protein alignment was performed as follows: A phylogenetic tree was reconstructed using a Neighbour Joining (NJ) approach as implemented in BioNJ 90 ; The likelihood of this topology was computed, allowing branch-length optimisation, using seven different models (JTT, LG, WAG, Blosum62, MtREV, VT and Dayhoff), as implemented in PhyML v3.0 91 . The two evolutionary models best fitting the data were determined by comparing the likelihood of the used models according to the AIC criterion 92 ; Then, ML trees were derived using these two models. All trees and alignments have been deposited in PhylomeDB 82 and can be browsed on-line (http://www.phylomedb.org, phylome codes 78 to 83). Trees were scanned to i) define orthology and paralogy relationships using a phylogeny-based, species overlap approach 28 ; ii) detect and date duplication events 29 , including large expansions of gene families; and iii) transfer functional annotations from one-to-one orthologs in S. cerevisiae. Unless indicated otherwise, all operations with phylogenetic trees were performed using scripts implemented within the ETE package 50 .

To reconstruct a species phylogeny, alignments of 603 proteins that had a single ortholog in all species considered in the phylome were concatenated into a single trimmed alignment of 288,995 positions. A Maximum Likelihood tree was reconstructed using phyML using the same parameters indicated for the trees in the phylome and the LG model. Branch support values were computed using the aLRT approach (see above) and based on an analysis of 100 bootstrap repetitions. In addition, a super-tree was reconstructed from the 4,965 gene phylogenies contained in the N. delphensis phylome, using a tree parsimony approach as implemented in DupTree 20 . Both approaches yielded identical topologies.

Average levels of sequence identity between orthologs were computed as follows: Each orthologous pair was aligned with MUSCLE v3.7 and the level of sequence identity was measured with trimAl V1.3 89 as the number of identical residues over the length of the shortest protein.

Substitution rate acceleration along specific lineages

In order to investigate how selective pressure varied along specific lineages in the phylogeny and whether positive selection was involved in the evolution and diversification of the Nakaseomyces, we used a subset of the original data. The subset analyzed included 7 species: the 6 Nakaseomyces and S. cerevisiae as outgroup. Following the same automated pipeline previously described to construct the Nakaseomyces phylome, we retrieved all the shared orthologous genes present in a single copy in all 7 genomes (i.e., one-to-one orthologs), we concatenated their respective alignments and estimated a phylogenetic tree using maximum likelihood. The resulting topology is consistent with that of the species tree in Figure 1. Using this phylogeny, we tested whether the rate of evolution of the one-to-one orthologs had accelerated along specific branches of the tree (i.e., affecting different species in the group), which would be consistent with either the relaxation of selective constraints or with the action of positive selection. We used the program codeml in the paml 4 package 93 to estimate the d N /d S rate ratio variation in each individual one-to-one ortholog, along particular branches in the tree. Subsequently, we compared pairs of nested models by means of a likelihood ratio test (LRT) where the degrees of freedom correspond to the difference in the number of parameters estimated in each model, and the distribution of the d N /d S rate is assumed to follow a chi square distribution. Values of dN/dS larger than 5 were filtered out.

We hypothesized an acceleration in gene evolution rate, and possible cases of positive selection, preceding the diversification of the ‘glabrata group’, that were perhaps involved in the capability of these species to become pathogenic. In LRT A (Additional file 12) we compare: i) A model that assumes an overall d N /d S rate ratio for all branches in the tree (omega 1) and a different rate for the branch that is ancestral to the ‘glabrata group’ (omega 2), where we hypothesize a rate acceleration; and ii) an alternative model that assumes one d N /d S ratio for the basal branches including S. cerevisiae, C. castellii and N. bacillisporus (omega 1), a different d N /d S rate in the branch ancestral to the ‘glabrata group’ (omega 2), and another d N /d S rate for the genus itself (omega 3). In this test, we were interested in verifying whether there was a significant increase in the d N /d S rate in the branch ancestral to the ‘glabrata group’ relative to the overall rate and whether this increase also occurred in the different species within the group.

To investigate whether pathogenic and non-pathogenic species were subjected to different selective pressure, we built two more LRTs to analyse another subset of species focusing on the ‘glabrata group’, we therefore excluded S. cerevisiae, and C. castellii, and used N. bacillisporus as the outgroup (Additional file 12). In LRT B we compared i) a model with two rates, one for the ‘glabrata group’, and one for the outgroup (N. bacillisporus); with ii) a model with four rates, one for the pathogenic species, one for the single non-pathogenic species in the subset (N. delphensis), one for the branch ancestral to the ‘glabrata group’, and one for the outgroup. In LRT C we compared i) a model with two rates, one for the ‘glabrata group’, and one for the outgroup (N. bacillisporus); with ii) a model with four rates, one for the the ‘glabrata group’, one for the branch ancestral to the ‘glabrata group’, one for the C. glabrata species itself and one for the outgroup.

Detection of adhesin genes

Putative adhesins were detected in the newly sequenced Nakaseomyces using a similarity search based on the known EPA genes in C. glabrata and the FLO genes in S. cerevisiae. Hits were filtered using the same thresholds applied during phylome reconstruction. Additionally, all proteomes were scanned for the presence of the Pfam domain PF10528, which is related to adhesins in fungi. The search was performed using HMMER v3, and proteins containing this domain with an e-value below 1e-05 were added to the selection of adhesins. Proteins were then scanned for undetermined regions as their sub-telomeric location can cause problems in the assembly. Proteins containing more than 33% of undetermined regions were excluded from further analysis (this affected 8 proteins, of which 1 in C. castellii, 2 in N. delphensis, 4 in C. bracarensis, and 1 in C. nivariensis). Putative adhesins were then clustered using the TribeMCL 94 algorithm as implemented in scps (inflation = 1.5) 95 . Clusters were then used to infer phylogenetic trees. First the repetitive regions of each sequence were masked using SEG 96 . Alignments were then reconstructed using MUSCLE v3.7 84 followed by a maximum likelihood tree reconstruction as implemented in PhyML. The LG model was used along with four rate categories and invariant positions infered from the data.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

TG and CF analysed genome data and wrote the paper. AC, EP, JP, PW produced genome data. TM, PD produced tools for automatic genome annotation and analysis. OL produced tools for genome analysis and analysed genome data. SA, SB, LJ, AP, JR produced and analysed MAT cassettes data. RA, SA, LJ, CF produced PFGE and ploidy data. MBF, MMH, GA, RA, CB, SC, JAC, HD, AEA, JG, LM, CM, CN, EW, BD, CH analysed genome data. All authors read and approved the final manuscript.

Acknowledgements

We thank Brendan Cormack for sharing unpublished results about EPA genes and access to sequences. We thank our colleagues from the Genolevures network for their continued enthusiasm and support. TG’s research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP/2007-2013)/ERC Grant Agreement n.310325, a Grant from the Qatar National Research Fund grant (NPRP 5-298-3-086), and by a grant from the Spanish Ministry of Economy and Competitiveness (BIO2012-37161). CF’s research is funded in part by an “Attractivité” grant from the University Paris Sud. GA is a recipient of a Marie Curie grant (FP7-PEOPLE-2010-IEF-No.274223). SB, HD and RA are recipients of, respectively, a shared post-doctoral grant and a Ph.D. grant, from the Région Ile-de-France’s DIM Malinf program. JAC was supported by the Ph.D. Program in Computational Biology of the Instituto Gulbenkian de Ciência, Portugal (sponsored by Fundação Calouste Gulbenkian, Siemens SA, and Fundação para a Ciência e Tecnologia; SFRH/BD/33528/2008).

<p>Epidemiology of invasive candidiasis: a persistent public health problem</p>PfallerMADiekemaDJClin Microbiol Rev20072013316310.1128/CMR.00029-06179763717223626<p>Genome evolution in yeasts</p>DujonBShermanDFischerGDurrensPCasaregolaSLafontaineIDe MontignyJMarckCNeuvegliseCTallaENature2004430354410.1038/nature0257915229592<p>Phylogenetic circumscription of <it>Saccharomyces</it>, <it>Kluyveromyces</it> and other members of the <it>Saccharomycetaceae</it>, and the proposal of the new genera <it>Lachancea</it>, <it>Nakaseomyces</it>, <it>Naumovia</it>, <it>Vanderwaltozyma</it> and <it>Zygotorulaspora</it></p>KurtzmanCPFEMS Yeast Res2003423324510.1016/S1567-1356(03)00175-214654427<p>Phenotypic and molecular characterization of <it>Candida nivariensis</it> sp. nov., a possible new opportunistic fungus</p>Alcoba-FlorezJMendez-AlvarezSCanoJGuarroJPerez-RothEdel Pilar ArevaloMJ Clin Microbiol2005434107411110.1128/JCM.43.8.4107-4111.2005123398616081957<p><it>Candida bracarensis</it> sp. nov., a novel anamorphic yeast species phenotypically similar to <it>Candida glabrata</it></p>CorreiaASampaioPJamesSPaisCInt J Syst Evol Microbiol200656Pt 131331716403904<p><it>Candida nivariensis</it>, an emerging pathogenic fungus with multidrug resistance to antifungal agents</p>BormanAMPetchRLintonCJPalmerMDBridgePDJohnsonEMJ Clin Microbiol20084693393810.1128/JCM.02116-07226833218199788<p><it>Candida bracarensis</it> detected among isolates of <it>Candida glabrata</it> by peptide nucleic acid fluorescence in situ hybridization: susceptibility data and documentation of presumed infection</p>BishopJAChaseNMagillSSKurtzmanCPFiandacaMJMerzWGJ Clin Microbiol20084644344610.1128/JCM.01986-07223811418077641<p>Yeast-like fungi of the human intestinal tract</p>AndersonHWJ Infect Dis19172134138510.1093/infdis/21.4.341<p>Biogeography of the yeasts of ephemeral flowers and their insects</p>LachanceMAStarmerWTRosaCABowlesJMBarkerJSJanzenDHFEMS Yeast Res2001B18<p><it>Torulopsis castellii</it> sp.nov. a yeast isolated from a Finnish soil</p>CapriottiAJ Gen Microbiol196126414310.1099/00221287-26-1-4113876389<p>A site-specific endonuclease essential for mating-type switching in <it>Saccharomyces cerevisiae</it></p>KostrikenRStrathernJNKlarAJHicksJBHeffronFCell19833516717410.1016/0092-8674(83)90219-26313222<p>Evolution of the <it>MAT</it> locus and its Ho endonuclease in yeast species</p>ButlerGKennyCFaganAKurischkoCGaillardinCWolfeKHProc Natl Acad Sci U S A20041011632163710.1073/pnas.030417010134179914745027<p>From <it>Saccharomyces cerevisiae</it> to <it>Candida glabrata</it> in a few easy steps: important adaptations for an opportunistic pathogen</p>RoetzerAGabaldonTSchullerCFEMS Microbiol Lett20113141910.1111/j.1574-6968.2010.02102.x301506420846362<p>Evolution of pathogenicity and sexual reproduction in eight <it>Candida</it> genomes</p>ButlerGRasmussenMDLinMFSantosMASakthikumarSMunroCARheinbayEGrabherrMForcheAReedyJLNature200945965766210.1038/nature08064283426419465905<p>Complete mitochondrial genome sequences of three <it>Nakaseomyces</it> species reveal invasion by palindromic GC clusters and considerable size expansion</p>BouchierCMaLCrenoSDujonBFairheadCFEMS Yeast Res200991283129210.1111/j.1567-1364.2009.00551.x19758332<p>Mechanisms of chromosome number evolution in yeast</p>GordonJLByrneKPWolfeKHPLoS Genet20117e100219010.1371/journal.pgen.1002190314100921811419<p>Analysis of centromere function in <it>Saccharomyces cerevisiae</it> using synthetic centromere mutants</p>MurphyMRFowlkesDMFitzgerald-HayesMChromosoma199110118919710.1007/BF003553681790732<p>A surprisingly large RNase P RNA in <it>Candida glabrata</it></p>KachouriRStribinskisVZhuYRamosKSWesthofELiYRNA2005111064107210.1261/rna.2130705137079115987816<p>Large telomerase RNA, telomere length heterogeneity and escape from senescence in <it>Candida glabrata</it></p>Kachouri-LafondRDujonBGilsonEWesthofEFairheadCTeixeiraMTFEBS Lett20095833605361010.1016/j.febslet.2009.10.03419840797<p>DupTree: a program for large-scale phylogenetic analyses using gene tree parsimony</p>WeheABansalMSBurleighJGEulensteinOBioinformatics2008241540154110.1093/bioinformatics/btn23018474508<p>The tree versus the forest: the fungal tree of life and the topological diversity within the yeast phylome</p>Marcet-HoubenMGabaldonTPLoS One20094e435710.1371/journal.pone.0004357262981419190756<p>Three mating type-like loci in <it>Candida glabrata</it></p>SrikanthaTLachkeSASollDREuk Cell2003232834010.1128/EC.2.2.328-340.2003<p>Evidence from comparative genomics for a complete sexual cycle in the ‘asexual’ pathogenic yeast <it>Candida glabrata</it></p>WongSFaresMAZimmermannWButlerGWolfeKHGenome Biol20034R1010.1186/gb-2003-4-2-r1015130012620120<p>Comparative genomics in hemiascomycete yeasts: evolution of sex, silencing, and subtelomeres</p>FabreEMullerHTherizolsPLafontaineIDujonBFairheadCMol Biol Evol20052285687310.1093/molbev/msi07015616141KurtzmanCPFellJWThe Yeasts, A taxonomic StudyLondon: Elsevier Science41998<p>Evolutionary erosion of yeast sex chromosomes by mating-type switching accidents</p>GordonJLArmisénDProux-WéraEÓhÉigeartaighSSByrneKPWolfeKHProc Natl Acad Sci U S A2011108200242002910.1073/pnas.1112808108325016922123960<p>The intronome of budding yeasts</p>NeuvegliseCMarckCGaillardinCC R Biol201133466267010.1016/j.crvi.2011.05.01521819948<p>Large-scale assignment of orthology: back to phylogenetics?</p>GabaldonTGenome Biol2008923510.1186/gb-2008-9-10-235276086518983710<p>Assigning duplication events to relative temporal scales in genome-wide studies</p>Huerta-CepasJGabaldonTBioinformatics201127384510.1093/bioinformatics/btq60921075746<p>Robustness–it’s not where you think it is</p>WolfeKNature Genet2000253410.1038/7556010802639<p>The Yeast Gene Order Browser: combining curated homology and syntenic context reveals gene fate in polyploid species</p>ByrneKPWolfeKHGenome Res2005151456146110.1101/gr.3672305124009016169922<p>Increased glycolytic flux as an outcome of whole-genome duplication in yeast</p>ConantGCWolfeKHMolec Systems Biol20073129<p>Fermentative lifestyle in yeasts belonging to the <it>Saccharomyces</it> complex</p>MericoASuloPPiskurJCompagnoCFEBS J200727497698910.1111/j.1742-4658.2007.05645.x17239085<p>Transition of the ability to generate petites in the <it>Saccharomyces</it>/<it>Kluyveromyces</it> complex</p>FeketeVCiernaMPolakovaSPiskurJSuloPFEMS Yeast Res200771237124710.1111/j.1567-1364.2007.00287.x17662054<p><it>Candida glabrata PHO4</it> is necessary and sufficient for Pho2-independent transcription of phosphate starvation genes</p>KerwinCLWykoffDDGenetics200918247147910.1534/genetics.109.101063269175619332882<p>Novel acid phosphatase in <it>Candida glabrata</it> suggests selective pressure and niche specialization in the phosphate signal transduction pathway</p>OrkwisBRDaviesDLKerwinCLSanglardDWykoffDDGenetics201018688589510.1534/genetics.110.120824297528920739710<p><it>Saccharomyces cerevisiae PAU</it> genes are induced by anaerobiosis</p>RachidiNMartinezMJBarrePBlondinBMol Microbiol2000351421143010760143<p>Nicotinic acid limitation regulates silencing of <it>Candida</it> adhesins during UTI</p>DomergueRCastanoIDe Las PenasAZupancicMLockatellVHebelJRJohnsonDCormackBPScience200530886687010.1126/science.110864015774723<p>Why do some yeast species require niacin for growth? Different modes of NAD synthesis</p>LiYFBaoWGFEMS Yeast Res2007765766410.1111/j.1567-1364.2007.00231.x17425674<p>Birth of a metabolic gene cluster in yeast by adaptive gene relocation</p>WongSWolfeKHNature Genet20053777778210.1038/ng158415951822<p>Analysis of the seven-member <it>AAD</it> gene set demonstrates that genetic redundancy in yeast may be more apparent than real</p>DelneriDGardnerDCOliverSGGenetics199915315911600146087010581269<p>Molecular and physiological aspects of alcohol dehydrogenases in the ethanol metabolism of <it>Saccharomyces cerevisiae</it></p>de SmidtOdu PreezJCAlbertynJFEMS Yeast Res201212334710.1111/j.1567-1364.2011.00760.x22094012<p>The omega-site sequence of glycosylphosphatidylinositol-anchored proteins in <it>Saccharomyces cerevisiae</it> can determine distribution between the membrane and the cell wall</p>FriemanMBCormackBPMol Microbiol20035088389610.1046/j.1365-2958.2003.03722.x14617149<p>An adhesin of the yeast pathogen <it>Candida glabrata</it> mediating adherence to human epithelial cells</p>CormackBPGhoriNFalkowSScience199928557858210.1126/science.285.5427.57810417386<p>Virulence-related surface glycoproteins in the yeast pathogen <it>Candida glabrata</it> are encoded in subtelomeric clusters and subject to <it>RAP1</it>- and <it>SIR</it>-dependent transcriptional silencing</p>De Las PenasAPanSJCastanoIAlderJCreggRCormackBPGenes & Dev2003172245225810.1101/gad.112100324040667<p>Sweet-talk: role of host glycosylation in bacterial pathogenesis of the gastrointestinal tract</p>MoranAPGuptaAJoshiLGut2011601412142510.1136/gut.2010.21270421228430<p>The cell wall of the human pathogen <it>Candida glabrata</it>: differential incorporation of novel adhesin-like wall proteins</p>de GrootPWKraneveldEAYinQYDekkerHLGrossUCrielaardWde KosterCGBaderOKlisFMWeigMEuk Cell200871951196410.1128/EC.00284-08<p><it>Candida glabrata</it> Pwp7p and Aed1p are required for adherence to human endothelial cells</p>DesaiCMavrianosJChauhanNFEMS Yeast Res20111159560110.1111/j.1567-1364.2011.00743.x320204221726406<p>Glycan microarray analysis of <it>Candida glabrata</it> adhesin ligand specificity</p>ZupancicMLFriemanMSmithDAlvarezRACummingsRDCormackBPMol Microbiol20086854755910.1111/j.1365-2958.2008.06184.x18394144<p>ETE: a python Environment for Tree Exploration</p>Huerta-CepasJDopazoJGabaldonTBMC Bioinforma2010112410.1186/1471-2105-11-24<p>The epithelial adhesin 1 (Epa1p) from the human-pathogenic yeast Candida glabrata: structural and functional study of the carbohydrate-binding domain</p>IelasiFSDecanniereKWillaertRGActa Crystallogr201268210217<p>Evolutionary rate variation among vertebrate beta globin genes: implications for dating gene family duplication events</p>AguiletaGBielawskiJPYangZGene2006380212910.1016/j.gene.2006.04.01916843621<p>Primary structure and transcription of an amplified genetic locus: the <it>CUP1</it> locus of yeast</p>KarinMNajarianRHaslingerAValenzuelaPWelchJFogelSProc Natl Acad Sci U S A19848133734110.1073/pnas.81.2.3373446716364141<p>Genomic polymorphism in the population of <it>Candida glabrata</it>: gene copy-number variation and chromosomal translocations</p>MullerHThierryACoppeeJYGouyetteCHennequinCSismeiroOTallaEDujonBFairheadCFungal Genet Biol20094626427610.1016/j.fgb.2008.11.00619084610<p>A family of glycosylphosphatidylinositol-linked aspartyl proteases is required for virulence of <it>Candida glabrata</it></p>KaurRMaBCormackBPProc Natl Acad Sci U S A20071047628763310.1073/pnas.0611195104186350417456602<p>Genetic variation of the repeated <it>MAL</it> loci in natural populations of <it>Saccharomyces cerevisiae</it> and <it>Saccharomyces paradoxus</it></p>NaumovGINaumovaESMichelsCAGenetics199413680381212058868005435<p>Genetic mapping of the alpha-galactosidase <it>MEL</it> gene family on right and left telomeres of <it>Saccharomyces cerevisiae</it></p>NaumovGINaumovaESLouisEJYeast19951148148310.1002/yea.3201105127597853<p>Lager yeasts possess dynamic genomes that undergo rearrangements and gene amplification in response to stress</p>JamesTCUsherJCampbellSBondUCurrent Genet20085313915210.1007/s00294-007-0172-8<p>Discovering the secrets of the Candida albicans agglutinin-like sequence (ALS) gene family–a sticky pursuit</p>HoyerLLGreenCBOhSHZhaoXMed Mycol20084611510.1080/13693780701435317274288317852717<p>Adaptive evolution of highly mutable loci in pathogenic bacteria</p>MoxonERRaineyPBNowakMALenskiRECurr Biol19944243310.1016/S0960-9822(00)00005-17922307<p>Acquisition of prokaryotic genes by fungal genomes</p>Marcet-HoubenMGabaldonTTrends Genet2010265810.1016/j.tig.2009.11.00719969385<p>Saccharomyces delphensis nov. spec.; a new yeast from South African dried figs</p>Van Der WaltJPTscheuschnerITAntonie Van Leeuwenhoek19562216216610.1007/BF0253832313340702<p><it>Kluyveromyces bacillisporus</it> sp. nov., a yeast from Emory Oak exudate</p>LachanceMAPhaffHJStarmerWTInt J Syst Bact19931115119<p>Gas analyses of anaerobic or microaerophillic generating systems using Gas Chromatography</p>RuangrungroteSIntasornASindermsukJJ Metals and Minerals2008181316<p>Long range control circuits within mitochondria and between nucleus and mitochondria. I. Methodology and phenomenology of suppressors</p>DujardinGPajotPGroudinskyOSlonimskiPPMol Gen Genet198017946948210.1007/BF002717367003299<p>Methods for studying the genetics of mitochondria</p>HauswirthWWLimLODujonBTurnerGMitochondria: A Pratical ApproachOxford: IRL PressDarley-Usmar VM, Rickwood D, Wilson MT1987171244<p>High quality draft sequences for prokaryotic genomes using a mix of new sequencing technologies</p>AuryJMCruaudCBarbeVRogierOMangenotSSamsonGPoulainJAnthouardVScarpelliCArtiguenaveFBMC Genomics2008960310.1186/1471-2164-9-603262537119087275<p>Conrad: gene prediction using conditional random fields</p>DeCaprioDVinsonJPPearsonMDMontgomeryPDohertyMGalaganJEGenome Res2007171389139810.1101/gr.6558107195090717690204<p>AUGUSTUS: ab initio prediction of alternative transcripts</p>StankeMKellerOGunduzIHayesAWaackSMorgensternBNucleic Acids Res200634Web Server issueW435W439153882216845043<p>EMBOSS: the European Molecular Biology Open Software Suite</p>RicePLongdenIBleasbyATrends Genet20001627627710.1016/S0168-9525(00)02024-210827456<p>Gene finding in novel genomes</p>KorfIBMC Bioinforma200455910.1186/1471-2105-5-59<p>Gene prediction in novel fungal genomes using an ab initio algorithm with unsupervised training</p>Ter-HovhannisyanVLomsadzeAChernoffYOBorodovskyMGenome Res2008181979199010.1101/gr.081612.108259357718757608<p>Using Geneid to Identify Genes</p>BlancoEParraGGuigoRCurrent Protocols in BioinformaticsNew York: John Wiley and SonsBaxevanis A2002<p>JIGSAW: integration of multiple sources of evidence for gene prediction</p>AllenJESalzbergSLBioinformatics2005213596360310.1093/bioinformatics/bti60916076884<p>Phylogenetic and structural analysis of centromeric DNA and kinetochore proteins</p>MeraldiPMcAinshADRheinbayESorgerPKGenome Biol20067R2310.1186/gb-2006-7-3-r23155775916563186<p>tRNomics: analysis of tRNA genes from 50 genomes of Eukarya, Archaea, and Bacteria reveals anticodon-sparing strategies and domain-specific features</p>MarckCGrosjeanHRNA200281189123210.1017/S1355838202022021137033212403461<p>tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence</p>LoweTMEddySRNucleic Acids Res19972559559641465259023104<p>Basic local alignment search tool</p>AltschulSFGishWMillerWMyersEWLipmanDJJ Mol Biol19902154034102231712<p>Infernal 1.0: inference of RNA alignments</p>NawrockiEPKolbeDLEddySRBioinformatics2009251335133710.1093/bioinformatics/btp157273231219307242<p>Genolevures: protein families and synteny among complete hemiascomycetous yeast proteomes and genomes</p>ShermanDJMartinTNikolskiMCaylaCSoucietJLDurrensPNucleic Acids Res200937Database issueD550D554268650419015150<p>Annotating Non-Coding RNAs with Rfam</p>Griffiths-JonesSCurr Protoc Bioinformatics200512.5.112.5.12<p>PhylomeDB v3.0: an expanding repository of genome-wide collections of trees, alignments and phylogeny-based orthology and paralogy predictions</p>Huerta-CepasJCapella-GutierrezSPryszczLPDenisovIKormesDMarcet-HoubenMGabaldonTNucleic Acids Res201139Database issueD556D560301370121075798<p>Identification of common molecular subsequences</p>SmithTFWatermanMSJ Mol Biol198114719519710.1016/0022-2836(81)90087-57265238<p>MUSCLE: multiple sequence alignment with high accuracy and high throughput</p>EdgarRCNucleic Acids Res2004321792179710.1093/nar/gkh34039033715034147<p>Recent developments in the MAFFT multiple sequence alignment program</p>KatohKTohHBrief Bioinfo2008928629810.1093/bib/bbn013<p>DIALIGN-TX: greedy and progressive approaches for segment-based multiple sequence alignment</p>SubramanianARKaufmannMMorgensternBAlgorithms Mol Biol20083610.1186/1748-7188-3-6243096518505568<p>Heads or tails: a simple reliability check for multiple sequence alignments</p>LandanGGraurDMol Biol Evol2007241380138310.1093/molbev/msm06017387100<p>M-Coffee: combining multiple sequence alignment methods with T-Coffee</p>WallaceIMO’SullivanOHigginsDGNotredameCNucleic Acids Res2006341692169910.1093/nar/gkl091141091416556910<p>trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses</p>Capella-GutierrezSSilla-MartinezJMGabaldonTBioinformatics2009251972197310.1093/bioinformatics/btp348271234419505945<p>BIONJ: an improved version of the NJ algorithm based on a simple model of sequence data</p>GascuelOMol Biol Evol19971468569510.1093/oxfordjournals.molbev.a0258089254330<p> <b>Estimating maximum likelihood phylogenies with PhyML</b> </p>GuindonSDelsucFDufayardJFGascuelOMethods Mol Biol200953711313710.1007/978-1-59745-251-9_619378142<p>Information and an extension of the maximumlikehood principle</p>AkaikeHSecond International Symposium on Information TheoryBudapest (Hungary): Akademiai Kiado1973267281<p>PAML 4: phylogenetic analysis by maximum likelihood</p>YangZMol Biol Evol2007241586159110.1093/molbev/msm08817483113<p>An efficient algorithm for large-scale detection of protein families</p>EnrightAJVan DongenSOuzounisCANucleic Acids Res2002301575158410.1093/nar/30.7.157510183311917018<p>SCPS: a fast implementation of a spectral method for detecting protein families on a genome-wide scale</p>NepuszTSasidharanRPaccanaroABMC Bioinforma20101112010.1186/1471-2105-11-120<p>Statistics of local complexity in amino acid sequences and sequence databases</p>WoottonJCFederhenSComput Chem19931714916310.1016/0097-8485(93)85006-X