Skip to main content
Advertisement
  • Loading metrics

The complex evolution and genomic dynamics of mating-type loci in Cryptococcus and Kwoniella

  • Marco A. Coelho ,

    Contributed equally to this work with: Marco A. Coelho, Márcia David-Palma

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

    marco.dias.coelho@duke.edu (MAC); heitm001@duke.edu (JH)

    Affiliation Department of Molecular Genetics and Microbiology, Duke University Medical Center, Durham, North Carolina, United States of America

  • Márcia David-Palma ,

    Contributed equally to this work with: Marco A. Coelho, Márcia David-Palma

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

    Affiliation Department of Molecular Genetics and Microbiology, Duke University Medical Center, Durham, North Carolina, United States of America

  • Seonju Marincowitz,

    Roles Investigation, Writing – review & editing

    Affiliation Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa

  • Janneke Aylward,

    Roles Investigation, Writing – review & editing

    Affiliations Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa, Department of Conservation Ecology and Entomology, Stellenbosch University, Stellenbosch, South Africa

  • Nam Q. Pham,

    Roles Investigation

    Affiliation Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa

  • Andrey M. Yurkov,

    Roles Resources, Investigation, Writing – review & editing

    Affiliation Leibniz Institute DSMZ-German Collection of Microorganisms and Cell Cultures, Braunschweig, Germany

  • Brenda D. Wingfield,

    Roles Funding acquisition, Project administration, Resources, Writing – review & editing

    Affiliation Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa

  • Michael J. Wingfield,

    Roles Funding acquisition, Project administration, Resources, Writing – review & editing

    Affiliation Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa

  • Sheng Sun,

    Roles Investigation, Writing – review & editing

    Affiliation Department of Molecular Genetics and Microbiology, Duke University Medical Center, Durham, North Carolina, United States of America

  • Joseph Heitman

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

    marco.dias.coelho@duke.edu (MAC); heitm001@duke.edu (JH)

    Affiliation Department of Molecular Genetics and Microbiology, Duke University Medical Center, Durham, North Carolina, United States of America

Abstract

Sexual reproduction in basidiomycete fungi is governed by MAT loci (P/R and HD), which exhibit remarkable evolutionary plasticity, characterized by expansions, rearrangements, and gene losses often associated with mating system transitions. The sister genera Cryptococcus and Kwoniella provide a powerful framework for studying MAT loci evolution owing to their diverse reproductive strategies and distinct architectures, spanning bipolar and tetrapolar systems with either linked or unlinked MAT loci. Building on recent comparative genomic analyses, we generated additional chromosome-level assemblies, uncovering distinct trajectories shaping MAT loci organization. Contrasting with the small-scale expansions and gene acquisitions observed in Kwoniella, our analyses revealed independent expansions of the P/R locus in tetrapolar Cryptococcus, possibly driven by pheromone gene duplications. Notably, these expansions coincided with a pronounced GC-content reduction best explained by reduced GC-biased gene conversion following recombination suppression, rather than relaxed codon usage selection. Diverse modes of MAT locus linkage were also identified, including three previously unrecognized transitions: one resulting in a pseudobipolar arrangement and two leading to bipolarity. All three transitions involved translocations. In the pseudobipolar configuration, the P/R and HD loci remained on the same chromosome but genetically unlinked, whereas the bipolar transitions additionally featured rearrangements that fused the two loci into a nonrecombining region. Mating assays confirmed a sexual cycle in Cryptococcus decagattii, demonstrating its ability to undergo mating and sporulation. Progeny analysis in Kwoniella mangrovensis revealed substantial ploidy variation and aneuploidy, likely stemming from haploid–diploid mating, yet evidence of recombination and loss of heterozygosity indicates that meiotic exchange occurs despite irregular chromosome segregation. Our findings underscore the importance of continued diversity sampling and provide further evidence for convergent evolution of fused MAT loci in basidiomycetes, offering new insights into the genetic and chromosomal changes driving reproductive transitions.

Introduction

Sexual reproduction is a cornerstone of eukaryotic biology, generating genetic variation through recombination and allele segregation, which fuels adaptation in changing environments [13]. The Fungal Kingdom is amongst the most ecologically diverse, with species thriving in nearly every environment, acting as decomposers, symbionts, commensals, as well as pathogens of plants, animals, and other fungi [46]. This versatility is mirrored in a wide range of dispersal stages and reproductive strategies, from heterothallism, requiring genetically distinct individuals of opposite mating types, to homothallism, which allows self-fertilization by a single individual as well as outcrossing [7,8].

In heterothallic basidiomycete fungi, sexual compatibility is typically governed by two mating-type (MAT) loci. The pheromone/receptor (P/R) locus encodes at least one mating pheromone and one G-protein-coupled pheromone receptor, enabling mate recognition and fusion. The HD locus encodes two homeodomain transcription factors (HD1/Sxi1 and HD2/Sxi2) that heterodimerize post-mating to control dikaryotic growth and sexual development [9]. Successful mating occurs only between mating partners with different, compatible alleles of both MAT loci.

Most basidiomycetes exhibit a tetrapolar breeding system, with P/R and HD unlinked on separate chromosomes, allowing each meiotic event to generate up to four distinct mating types [10]. Some lineages, however, have evolved a bipolar system, in which sexual compatibility is controlled by a single MAT locus. This transition can occur either through loss of function of one locus in mating-type determination, as in Coprinellus disseminatus [11] and other mushroom-forming Agaricomycetes [1214], or via physical linkage and subsequent fusion (i.e., genetic linkage) of the P/R and HD loci. In the latter case, translocations or chromosome fusions bring the two loci together, followed by additional rearrangements that suppress recombination, resulting in a single, fused, nonrecombining MAT region that is often expanded, highly rearranged, and biallelic [1518]. Some basidiomycetes also exhibit a pseudobipolar configuration, where the two MAT loci lie on the same chromosome but remain genetically unlinked because they are sufficiently distant to allow recombination, as seen in skin-commensal/pathogenic Malassezia species [19,20]. Beyond heterothallism, several basidiomycetes reproduce without a partner. This strategy, broadly termed homothallism, encompasses diverse genetic mechanisms, including primary homothallism, where a single individual carries all MAT alleles necessary for sexual development [2124], and unisexual reproduction involving cells of the same mating type [25].

Evidence suggests that tetrapolarity is ancestral in basidiomycetes, with bipolarity via P/R-HD fusion arising independently multiple times [15,16,18,26,27]. This bipolar configuration, often observed in fungal species associated with plants or animals as commensals or pathogens, increases sibling compatibility from 25% (tetrapolar) to 50% (bipolar), thereby facilitating inbreeding [9,28,29]. This may benefit species colonizing hosts, where spore dispersal and access to unrelated mating partners may be more limited. Bipolar species with fused MAT loci are found across the three basidiomycete subphyla, including smut fungi of grasses [15,30,31], anther-smut Microbotryum infecting Caryophyllaceae [17,27,32,33], and two Tremellomycetes lineages: Trichosporonales [18] and the human-pathogenic Cryptococcus species [16,34,35].

An important aspect of fused MAT loci is the inclusion of genes beyond the core HD and P/R determinants, some encoding regulators of post-mating development that influence morphogenesis (e.g., STE20, required for polarity maintenance in dikaryotic hyphae [36]) and dispersal (e.g., SPO14, likely involved in spore production) [37]. Additionally, recombination suppression often extends beyond the core mating-type genes, encompassing larger regions of the MAT-containing chromosome [38]. While this may preserve beneficial allele combinations, it also contributes to genomic degeneration, including gene loss, accumulation of transposable elements (TEs), and extensive rearrangements. These dynamics parallel sex chromosome evolution in other eukaryotes, highlighting potential convergent evolution in sexual reproduction regulation [34,39].

The genus Cryptococcus includes both human-pathogenic species of critical importance and nonpathogenic saprobes. The pathogenic clade comprises Cryptococcus neoformans, C. deneoformans, and six species in the C. gattii complex [40,41], with C. neoformans ranking first on the World Health Organization’s list of critical fungal pathogens [42,43]. Nonpathogens include C. wingfieldii, C. amylolentus, C. floricola, C. depauperatus, and C. luteus [23,4447], along with newly identified lineages [48,49], many associated with insects, particularly bark beetles [50]. The sister genus, Kwoniella, consists of saprophytes found in diverse environments, including soil, seawater, plant material, and insect frass [46,48,49,51,52].

Sexual reproduction is well characterized in Cryptococcus and has been linked to pathogenicity in medically relevant species [5355]. A hallmark of the sexual cycle is the yeast-to-hypha transition, which occurs after mating between yeast cells of opposite mating types (a and α) or during unisexual reproduction [25]. This morphological switch aids nutrient foraging [56,57] and enhances survival against environmental stressors and microbial predators, thought to be major selective forces in Cryptococcus evolution [53,5862]. Hyphae eventually differentiate into basidia, where diploid nuclei undergo meiosis to produce basidiospores that are smaller than yeast cells and more resilient to environmental stresses. These spores facilitate alveolar deposition upon inhalation and subsequent dissemination within the host [6365]. In contrast, sexual reproduction in Kwoniella has been documented only in K. mangrovensis and K. heveanensis [52,66,67]. These species do not form aerial basidia with spore chains, likely making them less efficient at environmental dispersal than Cryptococcus.

Recent surveys have uncovered novel Cryptococcus and Kwoniella species [41,48,49,68,69], often from single isolates, limiting mating tests and hindering characterization of reproductive strategies and compatibility systems. In such cases, high-quality genome assemblies offer a powerful alternative, enabling comparative analyses of MAT loci to elucidate reproductive systems. While our recent work generated chromosome-level genome assemblies for several Cryptococcus and Kwoniella species, providing insights into their genomic architecture and evolution [49], this was restricted to single strains per species, precluding in-depth analyses of MAT locus variation.

Here, we examine a broader diversity of species and, where possible, strains of opposite mating types within the same species, enabling systematic characterization of MAT loci structure, chromosomal organization, and key evolutionary transitions. Our findings reveal tetrapolarity as the ancestral state of both Cryptococcus and Kwoniella, with P/R-HD loci fusion (i.e., genetic linkage) evolving independently three times: (i) in the common ancestor of pathogenic Cryptococcus species, (ii) in a recently identified nonpathogenic Cryptococcus species, and (iii) in one Kwoniella species. We further show that the P/R locus itself is highly dynamic, sharing only a core set of four genes across both genera, and has undergone lineage-specific expansions, particularly in Cryptococcus, where increased pheromone gene copy number appears to have driven further rearrangements. Experimental crosses also confirmed a previously unobserved sexual cycle in C. decagattii and provided the first direct evidence of recombination in K. mangrovensis, revealing meiotic exchange in progeny and nonhaploid offspring with potential for self-fertility. Collectively, our findings support that fused MAT loci have evolved multiple times independently in basidiomycetes, reinforcing the role of convergent evolution in shaping reproductive strategies and ecological adaptation.

Results

Chromosome-level assemblies and updated phylogenetic relationships of Cryptococcus and Kwoniella

To enable detailed analyses of MAT locus structure and evolution, we expanded our previous genomic dataset (29 assemblies; [49]) by incorporating genome sequences for 12 additional Cryptococcus and 7 Kwoniella strains, including opposite mating types where available. Among the Cryptococcus strains, four represent three currently undescribed species (Cryptococcus sp. 3, sp. 4, and sp. 5; Fig 1A), while the remaining eight are opposite mating types of pathogenic Cryptococcus species previously sequenced [49]. For Kwoniella, three strains belong to the recently described species K. ovata, K. endophytica, and K. fici [48,70,71], while the remaining four strains are different mating types of K. mangrovensis, K. europaea, K. botswanensis, and K. heveanensis, as previously determined through PCR and mating tests [52,66,67]. All but three genomes were assembled from long-read sequencing data (PacBio or Oxford Nanopore) and polished with short-read (Illumina) data, yielding chromosome-level genome assemblies. The exceptions, K. europaea PYCC6162, K. botswanensis CBS12717, and K. heveanensis BCC8398 were sequenced solely with Illumina, resulting in more fragmented assemblies. Analyzed strains were inferred to be haploid based on genome sizes consistent with those previously documented for haploid strains [49]. In total, the expanded dataset comprises 27 Cryptococcus strains (spanning 17 species) and 22 Kwoniella strains (spanning 18 species) (S1 Appendix).

thumbnail
Fig 1. Phylogeny, breeding systems, and MAT loci organization in Cryptococcus and Kwoniella.

(A) Maximum likelihood phylogeny of strains analyzed in this study, inferred from a concatenated data matrix comprising protein alignments of 3,086 single-copy genes. Clades are labeled as A–I, with clade E shown as a nested subclade within clade F for easier cross-referencing. Branch lengths are given as substitutions per site (scale bar). All branches are supported >95% by SH-aLTR and UFBoot tests unless otherwise indicated. Gene concordance factor (gCF) and site concordance factor (sCF) were assessed to evaluate genealogical concordance, revealing unresolved branching order of clades A, B, and C (dashed branches; see S1 Fig for details). A black asterisk denotes that VGV has not yet been associated with human infections, and a white asterisk marks strains sequenced solely with Illumina, yielding more fragmented assemblies. (B) Predicted breeding systems, MAT locus organization, and mating-type identity of the P/R locus. Breeding systems are classified as heterothallic or homothallic for species with defined sexual cycles, and as likely heterothallic for species inferred to exhibit heterothallism based on MAT gene content. (C) Heatmap of gene presence/absence and pseudogene annotations within inferred MAT loci. Genes are ordered by their frequency of presence across species, with the most conserved genes on the left. Genes marked as “unclear” indicate cases where presence within MAT is uncertain because locus boundaries could not be precisely determined. BSP3 and IKS1, evicted from the MAT locus in C. deneoformans [34], are listed as absent in this species. Gene essentiality, as predicted or experimentally validated in strain H99 in prior studies rather than direct assessments in each species, is displayed at the bottom. Genes scored as absent from MAT may still be present elsewhere in the genome. (D) Gene presence frequency plot summarizing panel C. Pseudogenes were treated as present for frequency calculations, while unclear cases were treated as absent. The identification of either SXI1 or SXI2 was scored as “present” and combined for frequency quantification. (E) MAT loci locations in Cryptococcus sp. 3 (CWM60451), K. fici, and K. europaea, illustrating bipolar and pseudobipolar arrangements, respectively. Genes typically found either within or near the HD and P/R loci in tetrapolar species are designated as HD- and P/R-associated genes, and their respective chromosomal locations are indicated by gold or teal vertical bars. The data underlying this figure can be found in S1 Appendix and at https://doi.org/10.5281/zenodo.16987438.

https://doi.org/10.1371/journal.pbio.3003417.g001

To establish evolutionary relationships across species and determine the placement of the newly sequenced taxa, we identified 3,086 single-copy orthologous genes shared across Cryptococcus and Kwoniella, along with three outgroup species (Tremella mesenterica, Saitozyma podzolica, and Bullera alba), and reconstructed phylogenies employing both Maximum Likelihood (ML) and coalescent-based methods. Both approaches produced largely congruent results, except for the variable placement of Cryptococcus clades B and C relative to clade A (Figs 1A and S1). This inconsistency, noted in earlier studies [23,45,47], remains unresolved with the current dataset. Despite this, phylogenetic placement of the newly identified Cryptococcus and Kwoniella species was resolved. Cryptococcus sp. 3, known from a single strain (CMW60451) isolated from a bark beetle (Lanurgus sp.) infesting twigs of the endangered conifer Widdringtonia cedarbergensis in the Cederberg Mountains of South Africa, was placed as the closest relative of C. depauperatus, albeit on a long branch. Cryptococcus sp. 4 [72] and Cryptococcus sp. 5 clustered with Cryptococcus sp. 6 (OR918) [49,73], together forming a sister clade to all other known Cryptococcus. Among Kwoniella, K. ovata and K. endophytica were identified as sister species closely related to K. dendrophila, while K. fici grouped with Kwoniella sp. 4. The new Cryptococcus and Kwoniella species will be formally described elsewhere.

Sexual reproduction, breeding systems, and MAT gene identification and organization across Cryptococcus and Kwoniella

Sexual cycles have been described for eight Cryptococcus species: seven primarily heterothallic (C. neoformans, C. deneoformans, C. gattii, C. bacillisporus, C. deuterogattii, C. amylolentus, C. floricola), and one homothallic (C. depauperatus) (Fig 1A) [23,47,7478]. In Kwoniella, sexual reproduction has been confirmed only in K. mangrovensis and K. heveanensis, both heterothallic [45,52,66]. We sought to identify sexual cycles in species where none had been described, restricting tests to cases where opposite mating types were available. Crosses were performed under conditions known to induce mating in Cryptococcus [79] and examined for the presence of hyphal growth or other sexual structures (basidia and basidiospores).

In C. decagattii, sexual reproduction was confirmed by crossing strains 7685027 (MATα) and CBS11687 (MATa), which produced sexual structures (S2A Fig). In contrast, crosses between C. tetragattii IND107 and CBS11718 did not yield sexual structures (S2B Fig). For Cryptococcus sp. 3, sister to C. depauperatus, we observed key differences: while C. depauperatus is homothallic and exhibits continuous hyphal growth [23], Cryptococcus sp. 3 grew mostly as yeast and did not produce sexual structures alone, suggesting it is unlikely to be homothallic (S2C Fig). Within clade D species, we tested Cryptococcus sp. 4, where two strains of opposite mating types are available. Although some hyphal growth was observed, it was inconsistent, and no discernible sexual structures formed even after prolonged incubation (S2D Fig). These negative results may reflect low mating efficiency under laboratory conditions, and future genetic approaches such as CRG1 deletion [141,142] could help uncover latent sexual potential in these lineages. Given the absence of confirmed sexual reproduction in most of the species tested, we resorted to genomic data to examine MAT gene content and organization, inferring breeding systems (heterothallic or homothallic) and classifying MAT loci configurations as tetrapolar, pseudobipolar, or bipolar, as summarized in Fig 1B.

Consistent with previous studies [34,78], we confirmed that all pathogenic clade A species, including C. decagattii and C. tetragattii, have a bipolar mating configuration supported by genome assemblies from both mating types. In contrast, most of the nonpathogenic Cryptococcus and Kwoniella species appeared to be heterothallic and tetrapolar, consistent with the proposed ancestral state of basidiomycetes [9,80,81]. Three exceptions were identified: Cryptococcus sp. 3 (clade C), K. europaea (clade E), and K. fici (clade H) (Fig 1B and 1E). In Cryptococcus sp. 3 and K. fici, BLAST searches revealed that the key MAT genes (HD and the pheromone receptor STE3) are close together on the same chromosome, consistent with bipolarity (Fig 1E). In contrast, K. europaea exhibited a pseudobipolar configuration, with HD and STE3 on the same chromosome but separated by ~8.95 Mb (Fig 1E). These findings prompted detailed characterization of MAT loci across species and reconstruction of the chromosomal changes underlying bipolar configurations.

Structure and evolution of MAT loci in Cryptococcus and Kwoniella

Identification and general organization of MAT loci.

The structure and gene composition of MAT loci have been described for several Cryptococcus species within clades A, B, and C [16,23,34,35,47,78,82], but in Kwoniella analyses have been limited to K. mangrovensis and K. heveanensis, based solely on fosmid libraries from a single mating type, leaving the precise structure of MAT loci unresolved [66,67]. To expand on these analyses, we combined genome-wide synteny comparisons with gene genealogies to delineate MAT loci across both genera. The HD locus was defined based on the presence of the HD1 and HD2 genes, while the P/R locus boundaries were delineated by identifying regions of disrupted synteny between mating types (see Materials and methods).

In tetrapolar Cryptococcus and Kwoniella species, the HD locus is compact, consisting only of the divergently transcribed HD1/SXI1 and HD2/SXI2 genes (S3 Fig). As in other basidiomycetes, mating-type differences are mainly confined to the N-terminal regions of the encoded proteins, and this two-gene structure is strongly conserved. This organization persists in Kwoniella species, even though the HD locus predominantly resides in subtelomeric regions, inherently more prone to rearrangements (S3JS3O Fig). By contrast, the P/R locus spans a larger chromosomal region and is more variable across species and between mating types of the same species (Figs 2 and S4). It is significantly expanded in tetrapolar Cryptococcus species, averaging 91.7 kb compared to 29.3 kb in Kwoniella (Figs 2 and S4; S2 Appendix; P < 0.0001, Mann–Whitney U test). Within Cryptococcus, sizes range from 84.5 kb in Cryptococcus sp. 5 (P/R a2) to 101.5 kb in C. floricola (P/R a1), while in Kwoniella, they range from 24.3 kb in K. newhampshirensis (P/R a1) to 34.4 kb in K. heveanensis (P/R a2) (S2 Appendix).

thumbnail
Fig 2. Synteny analysis of the P/R locus in tetrapolar Cryptococcus and Kwoniella species.

(A, B) Detailed views of the P/R locus of representative Cryptococcus species from clades B and D. In both clades, the P/R locus exhibits lower GC content compared to other genomic regions. In clade D species, the predicted P/R locus contains only a subset of the genes found within the P/R locus of clade B species or the MAT locus of pathogenic Cryptococcus species. The additional genes in clade B or pathogenic species are instead located in the immediate downstream region, suggesting independent expansion of this genomic segment in clade D species. (C, D) P/R locus region in representative Kwoniella species of clades E and G. In a2 strains, the STE12 gene appears truncated. Remnants of the NDC80 gene (in clade E/F species) or the BSP3 gene (in clade G species) are observed at the right edge of the P/R a2 allele, likely resulting from inversion events. In all panels, P/R-associated genes are colored teal if their orthologs in Cryptococcus pathogens are located within the MAT locus, with darker teal indicating genes positioned in the flanking regions, and bright green for genes present within the P/R locus of C. amylolentus, but absent in most Cryptococcus pathogens. The P/R allele of each strain (a1 or a2) is indicated on the left. Chromosomes inverted relative to their original assembly orientations are marked with asterisks. GC content is depicted as the deviation from the genome average, calculated in 0.5 kb nonoverlapping windows. Gene essentiality within and in the immediate vicinity of the P/R locus is inferred from predictions or experimental validations in C. neoformans H99 rather than direct assessments in these species. Other features are annotated as shown in the key. See S4 Fig for additional comparisons. The data underlying this figure can be found at https://doi.org/10.5281/zenodo.16987438.

https://doi.org/10.1371/journal.pbio.3003417.g002

Evolution of the P/R locus in Kwoniella.

Gene content analysis of the Kwoniella P/R locus revealed nine genes shared across all species (Figs 1C, 2C, and 2D; S3 Appendix). These include a single pheromone gene (MFα or MFa) containing the canonical CAAX motif at its C-terminus, a pheromone receptor (STE3α or STE3a), and two additional genes (STE20 and STE12) with established mating-related roles and found within MAT in Cryptococcus pathogens [34,83,84]. Other conserved genes include DBP10 and REI1, both implicated in ribosome biogenesis in Saccharomyces cerevisiae [85,86]; ATG7, an autophagy-related gene [87]; a C2 domain-containing protein (C2prot), whose ortholog in S. pombe (Ync13) coordinates exocytosis, endocytosis, and cell-wall integrity during cytokinesis [88]; and NDC80, encoding a component of the NDC80 kinetochore complex, essential for chromosome segregation and spindle checkpoint activity [89]. In K. heveanensis and other clade G species, two additional genes are present within the P/R locus: BPS3, encoding a protein of unknown function with BTP/PZ and CAP-Gly domains (also present in the MAT locus of Cryptococcus pathogens), and UPB6, a predicted deubiquitinating enzyme associated with the 26S proteasome [90]. Of these, DBP10, REI1, and NDC80 are predicted to be essential in S. cerevisiae and in C. neoformans, based on previous studies [85,86,89,91,92].

Across Kwoniella, the P/R a2 allele is generally larger than the a1 (Figs 2C, 2D, and S4; S2 Appendix), mainly due to the presence of an additional gene of unknown function. This size difference is also accompanied by truncation of STE12 within the a2 allele, with fragments flanking STE3 on both sides in several species (e.g., K. mangrovensis and K. pini; Figs 2C and S4). Fragments of the 3′ end of NDC80 and BSP3 are likewise found at the right edge of the P/R a2 allele in certain clade E/F and clade G species, respectively, in addition to apparently intact copies (Figs 2C and S4). These fragments seem to be remnants of structural rearrangements, potentially associated with duplication tracks at inversion breakpoints. Reconstruction of the likely rearrangements leading to the extant P/R configuration in Kwoniella suggests three distinct inversions have shaped the locus in clades E/F: one involving STE3 and STE12, truncating STE12 in the a2 allele; a second relocating DBP10 and REI1 from the edge of the locus to its center; and a third moving NDC80 from the edge to the middle of the locus (S5A Fig). The first two inversions are conserved in clade G, indicating they are ancestral in Kwoniella, while clade G additionally underwent rearrangements repositioning BSP3-NDC80 and UBP6 to the center of the locus (S5B Fig). This suggests that the P/R a2 allele has experienced more modifications than the P/R a1 allele.

Additional complexity is seen in K. ovata and K. endophytica, where the P/R locus contains STE3a and MFa alleles alongside truncated versions of STE3α, with K. ovata additionally harboring a truncated MFα gene (Fig 3A). This mosaic configuration likely arose through intra-P/R recombination, as synteny analyses showed that the left side of the locus resembles that of P/R a2 strains, whereas the right side aligns more closely with the P/R a1 (Fig 3A). Gene genealogies further supported this scenario. While deep trans-specific polymorphism has been well-documented for STE3, STE12, and MF genes, reflecting their ancestral integration into the P/R locus across Cryptococcus and Kwoniella [23,47,67,78], shallower patterns were detected for NDC80, BSP3, ATG7, and UBP6. Specifically, NDC80 exhibited trans-specific polymorphism within clades E/F and G (Fig 3B), while ATG7, BSP3, and UBP6 displayed this pattern only within clade G (S5C Fig). Notably, the NDC80 allele at the right edge of the P/R locus in K. ovata and K. endophytica clustered more closely with a1 alleles from other species (Fig 3B), and a similar pattern was found in K. dendrophila, together pointing to a shared evolutionary history shaped by past intra-P/R recombination. This recombination event may have initially produced strains with intercompatible pheromone and receptor pairs, potentially enabling self-filamentous growth. However, subsequent loss of STE3α and MFα alleles appears to have eliminated self-filamentation. Consistently, solo cultures of K. ovata and K. endophytica on V8 pH 5 media did not form hyphae, indicating that extant species retain only relics of this ancestral recombination event. Collectively, these findings unveil a dynamic evolutionary history of the Kwoniella P/R locus, shaped by structural rearrangements, recombination, and small-scale lineage-specific expansions.

thumbnail
Fig 3. The P/R locus in K. ovata, K. endophytica, and K. dendrophila likely arose through intra-P/R recombination.

(A) Synteny analysis comparing P/R a1 and P/R a2 alleles in select Kwoniella species from clade F. The P/R loci of K. ovata and K. endophytica contain STE3a and MFa alleles, along with truncated (ψ) versions of STE3α. Additionally, K. ovata harbors a truncated MFα gene. Synteny analysis indicates that the left side of the P/R locus in K. ovata, K. endophytica, and K. dendrophila is structurally more similar to P/R a2 alleles of K. pini and K. shivajii, while the right side aligns more closely with the P/R a1 alleles of K. bestiolae and K. dejecticola. (B) Gene genealogy of the NDC80 gene inferred with IQ-TREE2 (model JTT + G4) showing trans-specific polymorphism across clade E/F species (labeled as 1) and within clade G species (labeled as 2). The NDC80 alleles of K. ovata, K. endophytica, and K. dendrophila cluster more closely with a1 alleles than with other a2 alleles. The data underlying this figure can be found at https://doi.org/10.5281/zenodo.16987438.

https://doi.org/10.1371/journal.pbio.3003417.g003

Evolution and expansion of the P/R locus in Cryptococcus.

Compared to Kwoniella, the P/R locus in tetrapolar Cryptococcus species is larger, averaging 95.4 kb in clade B and 86.9 kb in clade D, a difference not statistically significant (P = 0.063, Mann–Whitney U test; Figs 2 and S4; S2 Appendix). It also exhibits more complex rearrangements between mating types, which prevent accurate reconstruction of the evolutionary steps leading to the extant configurations (Figs 2, S4B, and S4E). Gene content analysis revealed striking variation between clades B and D (Fig 1C). Both share a core set of 7 genes (MF with two to four copies, STE3, STE12, BSP3, MYO2, STE20, and RPL39; S3 Appendix), but several genes found within the P/R locus in clade B species are instead located in the right flanking region of the locus in clade D species (Figs 2A, 2B, S4B, and S4E), and in Kwoniella these genes are similarly positioned just outside the locus (Figs 2C, 2D, S4H, S4K, and S4N). Despite only 7 genes being shared between clades B and D, the P/R locus is similar in size because clade D species harbor up to 15 distinct genes that are absent from the locus in other Cryptococcus or Kwoniella clades (Figs 1C, 2B, and S4E). Extending these comparisons across all tetrapolar species of both genera further revealed that only four genes (MF, STE3, STE20, and STE12) are conserved within the P/R locus, likely reflecting the ancestral state (Fig 1C and 1D; S3 Appendix).

Another notable distinction of the P/R locus in Cryptococcus is the presence of multiple pheromone genes per mating type [34,47]. Unlike Kwoniella, which retains a single pheromone gene per mating type, all heterothallic Cryptococcus species carry multiple copies, up to four per mating type. These genes are often arranged in divergent orientations (Figs 2A, 2B, S4B, and S4E), a configuration that may act as inverted repeats. Such an arrangement can facilitate inversion-loop formation during recombination, potentially contributing to structural instability and the expansion of the P/R locus boundaries in Cryptococcus. Together, these findings support a small ancestral P/R locus that expanded independently in descendant lineages by entrapping distinct gene sets through lineage-specific rearrangements.

A final striking observation was the consistently lower GC content of the P/R locus compared to genome-wide averages (Figs 2, S4A, and S4D; S4 Appendix). In C. amylolentus CBS6039 (clade B), the P/R locus has 51.24% GC versus 53.36% genome-wide (t test: t = 96.28, P < 0.0001; permutation test: P = 0.0303), with Z-scores ranging from −3.04 to −5.00 across clade B strains. This trend is even more pronounced in clade D species, where the difference can approach 10%. For example, in Cryptococcus sp. DSM108351, the P/R locus has 49.17% GC compared to a genome-wide average of 59.24% (t test: t = 216.92, P < 0.0001; permutation test: P = 0.0007), with Z-scores ranging from −7.56 to −6.71 across clade D strains. To test whether the GC depletion primarily reflects a neutral consequence of reduced GC-biased gene conversion (gBGC) [93,94] in a recombination-suppressed region or also involves relaxed selection on codon usage, we performed four complementary analyses (see S1 Text for details). All pointed to reduced gBGC as the main driver: (i) noncoding GC was significantly lower inside P/R than in matched same-chromosome windows, especially in clade D; (ii) coding sequences exhibited a marked shift toward AT-ending codons; (iii) codon-bias indices showed that P/R codon usage closely matched expectations based on nucleotide composition, unlike background genes; and (iv) codon adaptation index (CAI) differences were not significant after controlling for GC3 and gene length.

Independent origins and evolutionary outcomes of MAT locus fusion

Post-fusion structural dynamics of MAT in Cryptococcus pathogens.

The fusion of the P/R and HD loci into a single nonrecombining MAT locus is well established in clade A, which encompasses all human-pathogenic Cryptococcus species. We extended MAT locus analysis to all recognized species in this clade (Figs 1, 4, S6, and S7; S1 Appendix). For C. deneoformans, our dataset included congenic strains JEC21α and JEC20a, their progenitor strains NIH12α and NIH433a, and an additional MATa strain (NIH430) (S6 Fig and S1 Appendix). For C. neoformans, we examined multiple strains of both mating types across the four main lineages (VNI, VNII, VNBI, and VNBII), including eight newly sequenced genomes assembled telomere-to-telomere (T2T) (Fig 4A and S1 Appendix).

thumbnail
Fig 4. Synteny analysis of the mating-type locus (MATα and MATa) across different Cryptococcus neoformans lineages.

(A) Phylogenetic tree of representative C. neoformans strains from different VN lineages, inferred using maximum likelihood analysis of a concatenated matrix of protein alignments from 5,439 single-copy genes. Branch support is based on SH-aLTR and UFBoot tests. Mating types were determined based on the allelic version of the pheromone receptor gene (STE3). (B) Synteny analysis of chr. 5, where the MAT locus resides, showing significant conservation across strains, except within the MAT and centromere regions, which exhibit rearrangements. Strains are organized by mating type and lineage, as in panel D. (C) Box plot comparing MAT locus size between α and a strains of C. neoformans (C.n.) and C. gattii species complex (C.g.c.). The red dashed line, blue line, and boxes denote the mean value, median value, and interquartile range, respectively. Outliers are labeled with the corresponding strain name. Statistical significance (P-values) was determined by the Mann–Whitney U test. (D) Detailed synteny of the MAT locus, highlighting strain- and mating-type-specific rearrangements. The MATa configuration is more variable both across and within VN lineages, existing in three distinct configurations, whereas the MATα structure is more conserved, with limited rearrangements between ancestrally linked homeodomain (HD) and pheromone receptor (P/R) genes. Observed rearrangements appear to result from inversions mediated by identical copies of pheromone genes. KDZ transposons are detected within the MAT locus of certain strains. Genes typically associated with the P/R and HD loci in tetrapolar species are colored green and gold, respectively, with darker shades denoting genes flanking the MAT locus in Cryptococcus pathogens. Gene essentiality within the MAT locus, as predicted or experimentally validated in strain H99, is indicated by color-coded circles as given in the key. The data underlying this figure can be found at https://doi.org/10.5281/zenodo.16987438.

https://doi.org/10.1371/journal.pbio.3003417.g004

Synteny comparisons revealed that the MATa allele is generally longer than the MATα allele in both C. neoformans (P < 0.0001, Mann–Whitney U test) and the C. gattii species complex (P = 0.04142, Mann–Whitney U test; Fig 4C and S2 Appendix), with a similar trend in C. deneoformans, although limited sampling precluded statistical support. The MATa allele also exhibits greater structural diversity than MATα, as seen in the C. gattii complex (S7A Fig) and in C. neoformans, where frequent inversions were detected even among MATa strains of the same VN lineage (e.g., VNI strains Bt130 versus IUM96-2828, or strain Bt206 compared to other VNBII strains; Fig 4D). Inversion breakpoints were often associated with pheromone genes, consistent with the hypothesis that multiple pheromone copies may act as mediators of structural changes. In contrast, MATα is structurally more conserved with only limited inversions involving the pheromone–gene pairs flanking the PRT1 and ZNF1 genes (Figs 4D and S7). Comparison of the entire MAT chromosome across C. neoformans strains further showed that most of the intraspecific structural variation is confined to the MAT locus, with variation outside MAT largely restricted to the inherently dynamic centromeric and telomeric regions (Fig 4B).

As part of this analysis, we also detected significant structural divergence between the MATa alleles of C. neoformans VNBI (Ftc555-1 and Bt63) and VNI strains (Fig 4D). The VNI lineage has been hypothesized to have diverged initially with only the MATα allele, possibly due to a bottleneck associated with a small founding population, and later acquired the MATa allele via introgression from a VNBI MATa donor [95]. To date, only five VNI MATa strains have been identified [95100], three of which are now represented by high-quality genome assemblies (125.91 and Bt130, sequenced in this study, and IUM96-2828, sequenced previously [101] and annotated here; Fig 4D). This structural divergence did not contradict the introgression model but raised questions about how such differences could be reconciled with an introgressed origin.

To revisit this question with improved data, we reanalyzed SNP data leveraging newly generated reference genomes. Genome-wide SNP analysis, with VNBI MATa strain Ftc555-1 as the reference, revealed markedly lower SNP densities within the MATa locus in VNI strains compared to VNBII strains (Fig 5A and 5B). Conversely, across the rest of the genome, VNI MATa strains diverged more strongly from the VNBI reference, while VNBII strains showed lower SNP densities (Fig 5A and 5B). Notably, the typically high SNP density between VNI and VNBI strains was restored immediately downstream of MAT and ~5 genes upstream, consistent with recombination hotspots flanking the locus [102] and previously mapped introgression boundaries [95]. Phylogenetic analyses reinforced these patterns: VNBI strains clustered more closely with VNBII strains in trees inferred from genome-wide and chr. 5-specific SNPs (Fig 5C and 5D), whereas MAT-specific SNP trees placed VNI and VNBI MATa strains together, consistent with a more recent shared ancestry for the MATa allele (Fig 5E). Together, these findings refine the VNBI-to-VNI introgression model (Fig 5F) and indicate that lineage-specific rearrangements following introgression account for the structural differences observed.

thumbnail
Fig 5. SNP distribution, phylogenetic analysis, and revised introgression model of the MATa allele from VNBI into VNI.

(A) Genome-wide distribution of SNPs across chromosome 5 in C. neoformans MATa strains from different VN lineages relative to the C. neoformans Ftc555-1 VNBI genome reference. Empty spaces in the plots represent either sites identical to the reference genome (monomorphic sites) or regions excluded due to low coverage or repetitive/duplicated sequences. Strains on the y-axis are ordered by the total number of SNPs across the genome, with a color gradient (right) indicating SNP counts: darker colors represent higher SNP densities relative to the reference (strains positioned at the top), while lighter colors indicate fewer SNPs (strains at the bottom). Reads from the reference strain were also mapped as a control. The x-axis shows genomic coordinates. (B) Zoomed-in view of the MAT locus region plus 50 kb flanking regions, with genes and other features represented as in Fig 4. Notably, VNI strains exhibit a markedly lower number of SNPs at the MATa locus relative to the VNBI reference, compared to VNBII strains. (C) Midpoint-rooted tree based on genome-wide SNPs across all chromosomes, illustrating the evolutionary relationships of the strains. (D) Midpoint-rooted tree constructed using SNPs from chromosome 5 only, showing similar relationships to the genome-wide analysis. (E) Midpoint-rooted tree based on SNPs restricted to the MAT locus, highlighting a closer relationship between VNI and VNBI strains. In all trees, branches are shown as the number of substitutions per site (scale bars) and have >95% support based on SH-aLTR and UFBoot tests, unless otherwise indicated. “Ref” represents the reference genome assembly employed for read mapping and is thus identical to Ftc555-1 reads, serving as a control to validate the SNP calling pipeline. (F) Proposed model for the introgression of the VNBI MATa allele into VNI, shown alongside the key stages of the C. neoformans sexual cycle (right). Step 1: possible loss of the MATa allele in the VNI lineage, likely due to a bottleneck; step 2: mating and cell-cell fusion between a VNBI (MATa) and a VNI (MATα) strain; step 3: formation of dikaryotic hyphae, nuclear fusion, meiosis, recombination, and sporulation; step 4: some of the resulting F1 progeny emerged with a ~50:50 VNBI:VNI genome composition, retaining the VNBI MATa locus; step 5: repeated backcrossing of the MATa F1 progeny to VNI MATα strains progressively purging VNBI genomic regions while retaining the introgressed MATa allele; step 6: after ~10 backcrosses, the genome became ~99.95% VNI, with only the VNBI MATa locus and possibly a small fraction (~0.05%) of the VNBI genome remaining; step 7: over time, the introgressed VNBI MATa locus diverged and underwent rearrangements within the VNI background, reflecting strain-specific changes. The data underlying this figure can be found at https://doi.org/10.5281/zenodo.16987438.

https://doi.org/10.1371/journal.pbio.3003417.g005

In addition to introgression, other forces also shaped MAT architecture in C. neoformans. Species-specific gene losses (e.g., NCP1 and NCP2, which are absent in most species) and TE insertions contributed further to allele differentiation (S7 Fig). TE activity appears especially important, as the C. deneoformans MAT locus contains over five times more TEs than genome averages outside centromeres [16,34,103]. In line with this, we identified novel insertions of two distinct KDZ transposons, recently characterized in C. neoformans as large mobile elements (~11 kb) with terminal inverted repeats and target site duplications [104]. KDZx was present in strains 125.91, C45, PMHc1023, and T4, while KDZ1 was found in strain Ze90-1 (Fig 4D). Altogether, these findings underscore the dynamic evolution of the MAT locus in Cryptococcus pathogens, shaped by an interplay of structural rearrangements, gene loss, TE activity, and introgression.

Independent fusion of P/R and HD loci in Cryptococcus sp. 3.

Our initial screening of MAT gene content revealed that the P/R- and HD-associated genes are adjacent on the same chromosome in Cryptococcus sp. 3, indicating a fusion of these two regions (Fig 1E). While only a single strain (CMW60451) is available, limiting precise delineation of MAT boundaries, the minimum estimated size is ~90 kb based on the distance between the two most distal pheromone genes (Fig 6C). If boundaries are extended to include additional genes that are part of the MAT locus in pathogenic Cryptococcus, the locus could span ~120 kb, from RPL22 on the left to STE12 on the right (Fig 6C). Under this broader definition, 15 genes are shared between Cryptococcus sp. 3 and clade A species (Figs 1C and 6C; S3 Appendix), including four MFα pheromone gene copies and three genes typically associated with the HD locus (RPL22, CAP1, and SPO14). Additionally, two regions within the MAT locus contain TEs: one with a complete KDZ-like transposon and another with only remnants of this element (Fig 6C).

thumbnail
Fig 6. Transition to linked P/R-HD loci in Cryptococcus sp. 3 (CMW60451) and Cryptococcus pathogens.

(A) The karyotype of K. shandongensis (with 14 chrs.) served as the reference for reconstructing synteny blocks in pairwise comparisons. For simplicity, only synteny blocks corresponding to P/R- and HD-containing chromosomes (chrs. 12 and 5, colored teal and gold, respectively) were plotted in representative Cryptococcus species of clades A (C. neo. H99), B (C. amy. CBS6039), C (C. sp. CMW60451), and D (C. sp. DSM108351). Besides the MAT-containing chromosomes, synteny blocks of two additional chromosomes in K. shandongensis (chrs. 10 and 14), which correspond to a large portion of P/R and HD chromosomes in C. amylolentus were also plotted (color-coded light blue and green, respectively). Other synteny blocks can be visualized in S9A Fig. A red arrowhead pinpoints the predicted location of an inactivated centromere (ic) in Cryptococcus sp. CMW60451. (B) Linear chromosome plots depicting gene synteny conservation across species with zoomed-in views depicted in (C). In panel B, the MAT chromosome of bipolar species (C. neoformans and C. sp. 3, CMW60451) is shown twice: the left column depicts synteny relative to the P/R-containing chromosomes of tetrapolar species, and the right column depicts synteny relative to the HD-containing chromosomes.

https://doi.org/10.1371/journal.pbio.3003417.g006

Further analysis uncovered an unusual combination of MAT alleles. Unlike the bipolar Cryptococcus pathogens, where MATα strains carry HD1 (SXI1) linked to STE3α and MATa strains have HD2 (SXI2) linked to STE3a, Cryptococcus sp. 3 retains only SXI2 linked to STE3α. Although this distinct configuration suggested an independent fusion event at the origin of the MAT locus in Cryptococcus sp. 3, the unresolved phylogenetic placement of clade C relative to clade A left open the possibility that the two clades share a more recent common ancestor. In this scenario, the fusion of the P/R and HD loci could have occurred in their shared ancestor rather than independently in both lineages.

To test this hypothesis, we analyzed the phylogenetic clustering patterns of genes within the predicted MAT region shared with pathogenic Cryptococcus. Genes ancestrally present in the P/R locus, such as STE3, STE12, and MYO2, were excluded from the analysis because they exhibit deep trans-specific polymorphism predating the divergence of Cryptococcus species [23,47,67,78], making them uninformative. Instead, we focused on four genes (BSP3, PRT1, SPO14, and CAP1), which are part of MAT in pathogenic species and display mating-type-specific clustering within them [34]. The analysis revealed that Cryptococcus sp. 3 sequences do not group with either the a or α allele-specific clusters from pathogenic species; rather, they are distinct and fall outside these clusters, which is more consistent with an independent fusion event than one inherited from a shared ancestor (S8 Fig).

Having established from gene phylogenies that the MAT fusion in Cryptococcus sp. 3 likely arose independently, we next examined chromosome-scale structural changes underlying the P/R-HD linkage by reconstructing synteny blocks with SynChro [105]. K. shandongensis was selected as reference because (i) it is tetrapolar, with P/R and HD loci on separate chromosomes representing the ancestral state (Figs 1, S3N, and S4N); (ii) Kwoniella species show significantly fewer interchromosomal rearrangements than Cryptococcus, likely due to their simpler, smaller centromeres and lower TE density, as both high TE content and large, complex centromeres promote genomic rearrangements in Cryptococcus [49,106108]; and (iii) it retains a 14-chromosome karyotype, which is the predicted ancestral state for both genera [49] (Figs 6 and S9A).

The nuclear genome of Cryptococcus sp. 3 comprises 13 chromosomes (S1 Appendix), representing a reduction from the ancestral 14-chromosome karyotype. While we hypothesized that this reduction could explain the P/R-HD fusion, examination of the lost centromere showed it did not involve the MAT-containing chromosome (detailed in S2 Text and S9B Fig). Thus, physical linkage of the P/R and HD loci in Cryptococcus sp. 3 arose independently of the karyotypic reduction. We therefore compared the MAT-containing chromosome of Cryptococcus sp. 3 to the P/R- and HD-bearing chromosomes of K. shandongensis. Synteny analysis revealed significant alignment between chr. 11 of Cryptococcus sp. 3 and large regions of K. shandongensis chrs. 12 and 5, which carry P/R and HD, respectively (Fig 7A). Additionally, a small segment from the P/R chromosome is embedded towards the left end of chr. 11 within a region otherwise derived from the HD chromosome. These patterns suggest that the physical linkage likely began with a translocation that brought the two loci onto the same chromosome, initially far apart, followed by an inversion relocating the HD locus closer to the P/R locus, resulting in their fusion (events 1–3 in Fig 7A). Notably, the centromere-flanking regions of chr. 11 in Cryptococcus sp. 3 show considerable synteny to the CEN12-flanking region in the K. shandongensis P/R chromosome, indicating a shared evolutionary origin and suggesting that the translocation did not involve centromeric recombination but occurred elsewhere (Fig 7A).

thumbnail
Fig 7. Hypothesized mechanisms underlying the tetrapolar-to-bipolar transitions in Cryptococcus sp. 3 and in Cryptococcus pathogens.

Comparative synteny analysis between K. shandongensis and (A) Cryptococcus sp. 3 or (B) C. neoformans, illustrating the inferred sequence of chromosomal rearrangements leading to this transition. In both panels, the left side displays pairwise alignments, with syntenic regions connected by red (collinear) and gray (inverted) links, while the right side depicts a stepwise model of the proposed structural changes. The transition likely began with (1) an initial chromosomal break and reciprocal translocation (red dashed line), leading to (2) the repositioning of the P/R locus onto the same chromosome as the HD locus. (3) In Cryptococcus sp. 3, a subsequent inversion relocated the HD locus closer to the presumed ancestral position of the P/R locus, whereas in Cryptococcus pathogens, an inversion instead shifted the P/R locus toward the original location of the HD locus. (4) Additional structural modifications further refined the newly linked configuration, ultimately resulting in (5) the extant MAT locus organization observed in each species. These models suggest that despite independent evolutionary trajectories, both lineages underwent convergent genomic rearrangements that facilitated the transition to bipolar mating systems.

https://doi.org/10.1371/journal.pbio.3003417.g007

Revisiting the model of P/RHD loci linkage in Cryptococcus pathogens.

Prompted by these findings, we revisited the proposed model for the physical linkage and subsequent fusion of P/R and HD loci in Cryptococcus pathogens. This model, originally devised from comparisons between C. amylolentus (tetrapolar; clade B) and C. neoformans (bipolar; clade A), posited that the physical linkage was initiated by a chromosomal translocation mediated by ectopic recombination between repetitive elements within centromeric regions of P/R- and HD-containing chromosomes. This translocation would place the two MAT loci on the same chromosome, initially separated by the centromere. Subsequent rearrangements, such as inversions or transpositions, were hypothesized to bring the two loci closer together, ultimately resulting in their fusion and establishing the contiguous MAT locus of Cryptococcus pathogens [35,37].

To reassess this hypothesis, we compared multiple Cryptococcus species spanning the four clades, focusing on synteny and chromosomal architecture. In C. neoformans, the MAT locus resides on chr. 5. We found that its centromere (CEN5) shares conserved synteny on both flanking regions with CEN12 of K. shandongensis, as well as with CEN11 of Cryptococcus sp. 3 (clade C) and CEN9 of Cryptococcus sp. DSM108351 (clade D) (Fig 6B). These conserved synteny patterns indicate a shared evolutionary origin for these centromeres and strongly suggest that they were not directly involved in the translocation linking the P/R and HD loci in C. neoformans. In contrast, the P/R chromosome (chr. 10) of C. amylolentus (clade B) shows centromere-flanking regions that align with two distinct chromosomes, consistent with intercentromeric recombination and pointing to a derived rather than ancestral centromere structure.

Comparisons between C. neoformans and K. shandongensis further suggest a model in which the physical linkage and subsequent fusion of P/R and HD in C. neoformans arose through a mechanism similar to that proposed for Cryptococcus sp. 3, involving a chromosomal translocation outside the centromere followed by an inversion bringing the two loci into closer proximity (Fig 7B). In C. neoformans, however, the inversion seems to have relocated the P/R locus towards the HD locus near the left chromosomal end (event 3 in Fig 7B). These broader analyses refine the earlier model and suggest that the physical linkage of P/R and HD loci in Cryptococcus species likely arose through alternative mechanisms, independent of centromeric recombination.

Tetrapolar to bipolar and pseudobipolar transitions in Kwoniella.

We recently reported that karyotype reduction within Kwoniella occurred progressively and independently, often with formation of “giant” chromosomes (up to 18 Mb) through repeated fusions [49]. Although such events could potentially lead to P/R-HD physical linkage, only two Kwoniella species harbor both loci on the same chromosome (Fig 1B). In K. europaea (clade E), the HD and P/R loci are located on the same chromosome but remain far apart (~8.95 Mb) (Figs 1E and S4G), representing a pseudobipolar configuration. This arrangement likely arose from a lineage-specific rearrangement, in which the progenitor “giant” chr. 1 of clade E species underwent a translocation with chr. 2 [49]. Given their physical distance, the two MAT loci are still expected to recombine.

In K. fici, by contrast, the HD locus lies less than 13 kb from the P/R gene cluster, both located near the end of chr. 1 (Fig 8D). This “giant” chromosome likely arose independently through fusion events, similar to those documented in Kwoniella clades E and G [49]. After divergence from clade I (K. shandongensis/K. newhampshirensis), the chromosomes corresponding to chrs. 11 and 12 of K. shandongensis fused in the descendant lineage, producing the ancestral P/R chromosome, which persists as a single chromosome in most Kwoniella species (e.g., chr. 2 in Kwoniella sp. 4 DSM27419; Fig 8A) [49]. In K. fici, however, this ancestral fusion product fragmented further: one block now resides at the terminal region of K. fici chr. 1, while another segment lies on chr. 4 (Fig 8C). Similarly, most of the K. shandongensis/Kwoniella sp. 4 HD chromosome (chr. 5), aligns with the opposite end of K. fici chr. 1, while a small portion is on chr. 4, adjacent to the P/R-derived block. Only a few key HD-associated genes (SXI1, SXI2, and RPL22) are positioned near the P/R locus gene cluster at the right end of chr. 1 (Fig 8D). Notably, this derived bipolar arrangement retains both HD genes, unlike bipolar Cryptococcus species where opposite mating types carry only one. Overall, these observations indicate that complex translocations united blocks from ancestral P/R and HD chromosomes, producing the unique fused configuration in K. fici.

thumbnail
Fig 8. Linkage of P/R and HD loci in Kwoniella fici and MAT locus structure.

(A) The karyotype of K. shandongensis (with 14 chrs.) served as a reference for reconstructing synteny blocks in pairwise comparisons with Kwoniella sp. 4 and K. fici. (B) Phylogenetic relationships among selected Kwoniella species, highlighting their MAT locus organization and inferred breeding systems. (C) Synteny comparison between K. fici and its closest relative Kwoniella sp. 4, illustrating chromosomal rearrangements underlying the linkage of P/R and HD loci in K. fici. A red double-headed arrow indicates that the centromere-proximal regions of chr. 2 in Kwoniella sp. 4 correspond to a region near the fusion point on the K. fici “giant” chromosome (chr. 1), whereas the telomere-proximal regions of Kwoniella sp. 4 chr. 2 align with more internalized regions. This suggests that a large pericentric inversion targeting the centromere-adjacent region is associated with this fusion event, as previously reported for other Kwoniella species [49]. (D) Detailed gene-level organization of the MAT locus in K. fici compared to Kwoniella sp. 4. Each track (labeled a–d in dark circles) corresponds to a specific region marked by a black bar in panel C. Chromosomes inverted relative to their original assembly orientations are marked with asterisks.

https://doi.org/10.1371/journal.pbio.3003417.g008

Analysis of sexual reproduction in Kwoniella mangrovensis

Kwoniella mangrovensis (clade G) can reproduce sexually in the lab [52,66] and has only three chromosomes. To better understand its reproductive biology, we analyzed recombination, ploidy variation, and the potential for self-fertility in the progeny. The two originally sequenced strains, CBS8507 (P/R a1, HD b1) and CBS8886 (P/R a2, HD b1), share identical HD b1 alleles, making them incompatible for mating (Fig 9A). However, Guerreiro and colleagues [66] reported a third strain, CBS10435, as haploid and compatible with CBS8507. We sequenced CBS10435 with Illumina and confirmed it carries the P/R a2 allele and a distinct HD b2 allele, rendering it compatible with CBS8507 (Fig 9A).

thumbnail
Fig 9. Sexual reproduction and recombination in Kwoniella mangrovensis.

(A) Synteny analysis of the P/R and HD loci among three K. mangrovensis strains of different mating types (CBS8507, a1b1; CBS8886, a2b1; and CBS10435, a2b2). Note the pronounced divergence in the N-terminal domains of both HD1 (SXI1) and HD2 (SXI2) gene products, which are known to mediate mating-type specificity in other basidiomycetes. (B) Micrographs showing hyphal filaments extending from colony peripheries in a sexual cross between CBS8507 and CBS10435 after 2 weeks of incubation on CMA and V8 pH 5 media, in the dark, at room temperature. Insets depict two types of basidia: globose (a–c) and lageniform (b–d). Black arrows indicate clusters of cells emerging near basidia (either at the surface or embedded), scored as potential meiotic progeny (basidiospores). Single cells from these clusters were isolated using a micromanipulator, cultured into colonies, and genotyped by PCR-RFLP (see S5 Appendix). Putative recombinants were further analyzed using Illumina sequencing. Scale bars = 100 μm (25 μm in insets). (C) Genotypic analysis of selected meiotic progeny. SNP density between the parental strains was calculated as the number of SNPs per 10 kb (top). The genotypes of 9 segregants, derived from four distinct cell clusters, were inferred from SNP data and are depicted as follows: orange for regions inherited from CBS8507, purple for regions inherited from CBS10435, and brown for heterozygous regions (i.e., inherited from both parents). Instances of recombination or loss of heterozygosity (LOH) are detected by changes in genotype along the chromosomes. Discrepancies in ploidy, as inferred from FACS and sequencing read coverage, suggest potential genomic instability (marked by asterisks). Mating-type identity was inferred through sequencing, coverage analysis, and FACS. (D) Sequencing coverage plot for progeny MP59 with color-coded contributions from each parent. Haplotype blocks inferred from SNP data are overlaid for each chromosome for comparison (additional progeny data is presented in S8 Fig). (E) Self-filamentation phenotype of K. mangrovensis progeny. Parental strains CBS8507 and CBS10435 (grown individually and in co-culture) and their recovered progeny were cultivated on CMA at room temperature. Self-filamentation was assessed after 2 weeks of incubation. While neither parental strain exhibited self-filamentation in solo culture, their co-culture produced hyphal filaments and basidia. Progeny MP16, MP18, and MP54 also exhibited robust self-filamentation on both CMA and V8 pH 5 (see S9 Fig), whereas progeny MP49, MP51, and MP59 displayed weaker filamentous growth, detectable only on V8 pH 5 (S9 Fig). Scale bars = 200 μm. (F) Summary of the growth phenotype of the recovered progeny. The production of mycelium and basidia was classified as: extensive (+++), when observed across the entire periphery of the mating patch; moderate (++), when restricted to specific areas of the mating patch; poor (+), when limited to a single spot of the mating patch and slower to develop; and negative (−), when no filamentous growth was observed. Results represent observations from two independent tests.

https://doi.org/10.1371/journal.pbio.3003417.g009

Crosses between CBS8507 and CBS10435 on CMA and V8 pH 5 media produced basidia with characteristic globose and lageniform morphologies, consistent with prior descriptions (Fig 9B) [52]. Progeny were recovered by dissecting eight clusters of cells germinating from basidia along the hyphae (S5 Appendix). In total, 51 F1 progeny were obtained from both culture media. To avoid analyzing clonally derived genotypes, we screened for recombination with a PCR-RFLP approach using two markers per chromosome, positioned on either side of the centromere (S5 Appendix). Evidence of allele exchange and heterozygosity was detected, and nine F1 progeny with distinct genotypes were selected for whole-genome sequencing. SNPs between the parental strains served as markers to infer inheritance patterns associated with recombination and loss of heterozygosity (LOH), while read coverage analysis and FACS were combined to assess ploidy.

These analyses revealed that most F1 progeny were nonhaploid (Figs 9C, 9D, S10, and S11). Genotyping and FACS results were largely concordant, though some isolates (e.g., MP11 and MP18) showed discrepancies, suggesting variability within cell populations, possibly stemming from intrinsic genomic instability. Accordingly, mapping reads to a combined reference genome (CBS8507 and CBS10435) revealed uneven coverage for certain chromosomes. For example, CBS10435 chr. 2 exhibited low coverage in progeny MP59 (Fig 9D). Although CBS8507 and CBS10435 were previously reported as haploid [66], our FACS analysis showed that CBS10435 is predominantly diploid with a haploid subpopulation (Figs 9C and S11). This difference in parental ploidy likely underlies the aneuploidy observed in the progeny, as haploid–diploid matings are prone to chromosome mis-segregation, and the reduced karyotype of K. mangrovensis (3 chromosomes) may further amplify these effects.

Analysis of the MAT gene content revealed that some progeny carried fully compatible P/R and HD alleles (a1a2b1b2), potentially enabling self-fertility (Fig 9C). To investigate this, we cultured the F1 progeny under mating conditions and compared their phenotypes to a parental cross (Figs 9E, 9F, and S12). On CMA and V8 pH 5 media, progeny MP16, MP18, and MP54 with fully compatible MAT alleles produced hyphae and basidia similar to the parental cross. Some hyphal growth was also observed in solo culture of strains MP49, MP51, and MP59, exclusively on V8 pH 5, but the filaments were shorter and had no structures resembling basidia (S12 Fig). This morphology resembled that of previously described self-filamentous K. mangrovensis strains, which form irregular, highly branched hyphae without clamp cells [66]. Taken together, these analyses demonstrate that sexual reproduction in K. mangrovensis involves recombination and can generate nonhaploid progeny with potential for self-fertility.

Discussion

Our comparative genomic analysis revealed new insights into the evolution and structural dynamics of Cryptococcus and Kwoniella MAT loci, including distinct expansions of the P/R locus and additional independent HD-P/R fusion events, summarized in Fig 10 as a roadmap of key transitions.

thumbnail
Fig 10. Summary of the evolutionary transitions in MAT locus organization across Kwoniella and Cryptococcus species.

Schematic representation of key chromosomal rearrangements and evolutionary events underlying transitions from tetrapolar to bipolar and pseudobipolar mating configurations. The phylogenetic tree highlights inferred changes in MAT locus structure, with color-coded branches representing both extant and reconstructed MAT configurations. Dashed lines indicate unresolved phylogenetic relationships among clades A, B, and C. Insets summarize distinct evolutionary stages, including: the ancestral organization with unlinked and compact P/R and HD loci (1); small-scale expansion of the P/R locus in Kwoniella (2); two independent large-scale expansions of the P/R locus in Cryptococcus (5); relocation of P/R and HD loci onto the same chromosome in Kwoniella resulting in either pseudobipolar (3) or fused (4) MAT configurations; and independent HD-P/R fusion events with HD gene loss in Cryptococcus pathogens and Cryptococcus sp. 3 (6). These transitions illustrate the diverse pathways by which chromosomal rearrangements (CRs) have shaped MAT locus architecture in these fungal lineages.

https://doi.org/10.1371/journal.pbio.3003417.g010

Multiple independent P/R locus expansions in Cryptococcus and Kwoniella

The archetypal basidiomycete MAT loci are thought to consist of a compact P/R region with one pheromone receptor and one pheromone gene, and an unlinked HD region containing two divergently transcribed homeodomain transcription factors (HD1 and HD2). This configuration is retained in some lineages, such as in Malassezia species [20], but in most basidiomycetes the P/R locus has undergone expansions and structural changes, while the HD locus remains comparatively conserved (except in some rust fungi [109] and in mushroom-forming species with HD gene duplications [9,28]). For example, in Ustilago maydis, allele-specific differences in P/R locus size arise from two additional genes (lga2 and rga2) in the a2 allele that direct mitochondrial uniparental DNA inheritance [110,111]. Similarly, several Microbotryomycetes yeasts (Pucciniomycotina) harbor expanded P/R loci [26,112,113], as do tetrapolar Cryptococcus species [35,47,78]. Recent fully phased, chromosome-scale assemblies in rust fungi also uncovered multiple pheromone genes and STE3 receptor duplications on single haplotypes [109,114], though the full structure of the P/R locus in these fungi remains incompletely resolved.

Our analyses across tetrapolar Cryptococcus and Kwoniella support independent, lineage-specific P/R locus expansions. The inferred ancestral P/R locus likely comprised a conserved four-gene core (MF, STE3, STE12, and STE20) (Fig 10, event 1). In Kwoniella, this region has remained relatively compact, with 9 core conserved genes (Fig 10, event 2), though lineage-specific modifications were detected, including mating-type-specific changes (e.g., STE12 truncation in a2 alleles) and small-scale expansions through the inclusion of genes likely entrapped by local inversions. We note that these additions more likely reflect neutral consequences of inversions expanding recombination-suppressed regions, rather than adaptive recruitment, consistent with stepwise extension models of sex chromosome and MAT locus evolution [39,115]. In K. ovata, K. endophytica, and K. dendrophila, mosaic P/R configurations are better explained by past intra-locus recombination. This may have transiently enabled self-compatibility at the P/R locus, but the absence of self-filamentation in extant strains suggests a reversion to heterothallism. A full transition to homothallism would also require compatibility at the HD locus, either through mutations enabling pairing between otherwise nonself-compatible HD1/HD2 alleles or through bypass of HD function, as seen in C. depauperatus [23].

In contrast to Kwoniella, tetrapolar Cryptococcus species from clades B and D exhibit pronounced P/R locus expansion, but with limited gene-content overlap, indicating that these expansions occurred independently after divergence from a common ancestor (Fig 10, event 5). A distinctive feature of the Cryptococcus P/R locus is the duplication of pheromone (MF) genes, with each mating type typically carrying multiple copies arranged in divergent orientations. Similar patterns occur in Microbotryomycetes yeasts [26,112,113,116,117], and have also been reported in rust fungi [109,114]. However, in rusts, some of the duplicated MF genes encode distinct pheromones, possibly reflecting adaptation to different receptor specificities, whereas in Cryptococcus and Microbotryomycetes the MF copies encode identical mature peptides, suggesting functional redundancy rather than diversification.

Experimental evidence in C. deneoformans supports a dosage-based model, in which all three MFα gene copies are expressed and contribute additively to mating efficiency and filamentation [118]. Triple mutants show severe mating defects, while overexpression enhances differentiation, indicating that pheromone dosage promotes mating responsiveness [118]. Although direct functional evidence in other Cryptococcus species is lacking, these duplications may similarly enhance partner recognition. This interpretation aligns with sexual selection frameworks in fungi, where pheromone production can act as a costly, honest signal of mating quality (the handicap principle), which posits that only high-quality individuals can afford the cost of producing strong signals and are therefore more likely to be chosen as mates [119,120]. Studies in model yeasts have shown preferential mating toward high pheromone producers [121,122], but equivalent tests in basidiomycetes are limited, and the ecological context for such selection in Cryptococcus remains unclear. Why pheromone duplications recur in Cryptococcus but not in Kwoniella remains uncertain. Their absence in Kwoniella may reflect lineage-specific constraints or ecological differences—for instance, Kwoniella species produce embedded rather than aerial spores, which could reduce the need for long-range mate signaling—although the lack of ecological and population-level data prevents firm conclusions.

Notably, MF gene duplications can also introduce structural instability by facilitating the formation of inverted repeats, which can promote inversion loops and allele-specific structural rearrangements, potentially expanding recombination-suppressed regions between P/R alleles. One way to mitigate this instability is by clustering MF gene pairs more closely, thereby restricting the extent of rearrangements. Consistent with this, we observed reduced structural variation in C. neoformans MATα strains and in the C. gattii species complex, where the two MF gene pairs are positioned in close proximity, separated by only two genes. In contrast, MATa strains exhibit a more dispersed arrangement of MF genes, which appears to contribute to greater structural diversity across strains, including the C. neoformans VNI-specific rearrangements that arose following introgression of the MATa allele from the VNBI lineage. Once established, multiple MF gene copies may be maintained as templates for intrallelic repair via gene conversion, a nonreciprocal exchange that can occur during meiosis between homologous sequences but also at lower frequencies between sister chromatids, or within duplicated regions on a single chromosome during mitosis [123,124]. If confirmed, this mechanism would resemble the role of palindromic sequences in preserving gene integrity on the human Y chromosome [125127]. Introgression events, such as the transfer of MATa alleles between C. neoformans lineages, may further buffer against degeneration by resetting mutational load [128], thereby helping maintain complex MAT configurations.

Integrated with these structural features, our analyses revealed that the P/R locus in clades B and D Cryptococcus species has consistently lower GC% relative to genome-wide averages and that this depletion is best explained by reduced gBGC in recombination-suppressed regions rather than by relaxed selection on codon usage. gBGC normally favors the fixation of G/C alleles in regions of frequent recombination, and its reduction allows the AT-biased mutational spectrum to dominate over time, driving local base composition shifts. Similar GC depletion has been reported in the Trichosporonales MAT locus [18]. Interestingly, this trend is not observed in the MAT locus of pathogenic Cryptococcus, suggesting that genomic GC background also influences the outcome. In high-GC genomes (e.g., clades B and D), recombination suppression accentuates AT accumulation, whereas in lower-GC genomes (pathogenic Cryptococcus; [49]), the absence of a shift suggests they may already be near equilibrium under mutation pressure, masking localized effects.

Independent transitions to bipolar and pseudobipolar via chromosomal translocations and inversions

Beyond the previously documented tetrapolar-to-bipolar transition in pathogenic Cryptococcus lineages (clade A), we identified three additional transitions: two leading to bipolarity (Cryptococcus sp. 3 and K. fici; Fig 10, events 6 and 4), and one resulting in a pseudobipolar arrangement (K. europaea; Fig 10, event 3). Our findings suggest that the transition to bipolarity in Cryptococcus sp. 3 likely occurred independently of that in pathogenic species. First, fusion of the P/R and HD loci in Cryptococcus sp. 3 was unrelated to the reduction from 14 to 13 chromosomes and instead originated from a translocation followed by an inversion. Second, the MAT locus in this species retains only SXI2 (HD2) linked to the STE3α, unlike in pathogenic species where SXI2 is associated with STE3a. Third, phylogenetic analyses revealed that many genes within the Cryptococcus sp. 3 MAT locus do not group with a or α allele-specific clusters from pathogenic species. If the fusion had predated the divergence of these clades, trans-specific polymorphism would be expected, but this pattern was not observed, which is more consistent with an independent origin.

However, whether the initial translocation event that brought the P/R and HD loci onto the same chromosome occurred independently in Cryptococcus sp. 3 or in a common ancestor of clades A and C, remains unresolved. This uncertainty stems from three main factors. First, phylogenetic ties among clades A (bipolar), B (tetrapolar), and C (which includes Cryptococcus sp. 3 and C. depauperatus) are still ambiguous. Second, although our synteny analyses indicate that the translocation breakpoints at the origin of the physical linkage between MAT loci differ between Cryptococcus sp. 3 and pathogenic species (supporting independent events), these differences might also reflect subsequent rearrangements that accumulated along the two lineages. Third, we have previously reported that C. depauperatus also shows a partial merging of HD- and P/R-derived genes into the same genomic region, but with more extensive gene reshuffling and HD gene loss likely associated with a transition to homothallism [23]. Further taxon sampling, combined with phylogenomic approaches tailored to deep divergences, will be required to conclusively resolve these relationships and clarify the broader implications of these transitions.

The tetrapolar-to-bipolar transition in Cryptococcus pathogens has been hypothesized to result from a chromosomal translocation mediated by ectopic recombination between repetitive elements in the centromeric regions of the P/R- and HD-containing chromosomes [35]. This model had support from several observations: (i) Cryptococcus centromeres are enriched with long-terminal-repeat retrotransposons (LTRs) from the Ty3 and Ty1 superfamilies, which are shared among different centromeres and may facilitate recombination [23,47,49,129,130]; (ii) chromosomal arm exchanges occurred during the divergence of Cryptococcus lineages from their common ancestor [35,108]; and (iii) the repair of experimentally induced double-strand breaks at centromere-specific transposons in C. neoformans frequently results in chromosomal translocations [108]. However, our broader comparisons indicate that the centromere of the C. neoformans MAT chromosome, along with extended flanking regions, is conserved across most Cryptococcus species and in K. shandongensis, suggesting an ancestral organization.

Indeed, our analyses suggest that translocations outside centromeric regions may have played a more prominent role in repositioning the P/R and HD loci onto the same chromosome (Fig 7). Such noncentromeric translocations also appear to underlie the transition to bipolarity in several Ustilaginales species, including Sporisorium scitamineum [30] and in the common ancestor of Ustilago hordei, U. nuda, and U. bromivora [31]. Noncentromeric translocation events are also common across other genomic regions when comparing genomes of different Cryptococcus species [49] and even among conspecific strains of C. neoformans [101,131]. Whether such translocations are influenced by TEs has not yet been systematically analyzed, but studies in Saccharomyces cerevisiae have shown they can occur via homologous recombination between retrotransposons on different chromosomes [132]. Given the abundance of TEs in Cryptococcus genomes [49], they may likewise have facilitated the chromosomal relocation of MAT loci and the subsequent transition to bipolarity. Our findings offer an alternative to models invoking intercentromeric recombination as the primary event initiating the tetrapolar-to-bipolar transition in Cryptococcus pathogens.

Genetically linked P/R-HD loci can also arise from the fusion of entire ancestral P/R and HD chromosomes, as reported in Microbotryum violaceum on Silene paradoxa [27]. In Kwoniella, repeated and independent chromosome fusions have drastically reduced chromosome numbers, with some extant species retaining only 3 (clade E) or 5 (clades G and H) chromosomes from a 14-chromosome ancestral karyotype [49]. Yet, aside from the bipolar arrangement in K. fici (clade H) and the pseudobipolar arrangement in K. europaea (clade E), no other instances of P/R-HD linkage or colocation were detected in Kwoniella, and even in these species the transition reflects chromosomal translocations rather than whole-chromosome fusions.

In K. fici, the transition to fused MAT loci did not result in loss of SXI1 (HD1) or SXI2 (HD2), possibly suggesting a relatively recent fusion where insufficient time has passed for one HD gene to decay in each mating type, as observed in Cryptococcus sp. 3, Cryptococcus pathogens [133], and Trichosporonales [18]. Other known instances of HD gene loss in bipolar species, albeit distinct, include: (i) Microbotryum violaceum on Silene caroliniana, where HD1 is absent in a2 strains, while a1 retains both HD genes [27]; (ii) Microbotryum superbum, where loss of HD function in mating compatibility may have relaxed constraints on HD2 maintenance in a1 strains, leading to its disruption while a2 strains retain both HD genes [117]; and (iii) xerotolerant Wallemia species, where HD1 is retained in MAT a1, while the opposite mating type appears to lack any HD gene [134]. Because HD compatibility is no longer an independent determinant of mating success in bipolar species, maintaining both HD genes in each MAT allele may become dispensable, with loss potentially reflecting genomic streamlining once MAT loci are fused and recombination suppressed. Yet, as in K. fici, the retention of both HD genes in each allele remains the predominant organization in other bipolar basidiomycetes with fused MAT loci, including several Microbotryum [27], Sporisorium [30], and Ustilago species [31,135,136], possibly reflecting differences in mating strategies or genomic constraints. For instance, if both HD genes have additional roles beyond mating-type determination that do not require heterodimerization, their retention could be selectively maintained, although this possibility remains largely unexplored in basidiomycetes.

In contrast to K. fici, where the MAT loci are fused, K. europaea has the two loci ~8.5 Mb apart on the same chromosome. This distance makes tight genetic linkage unlikely, though extended recombination suppression over megabase-long spans has been reported in Neurospora tetrasperma and Microbotryum [39,137]. Because crosses in K. europaea have not yielded sexual reproduction [66], it remains unclear whether this configuration represents an incipient step toward bipolarity or a neutral byproduct of chromosomal rearrangements. If recombination persists, this system would functionally resemble tetrapolarity, as the two loci would still reassort in most meioses, making MAT loci colocation unlikely to meaningfully increase sibling compatibility. In this case, the arrangement may simply reflect a stochastic translocation rather than a selectively maintained feature.

Together, these observations reinforce the view that chromosomal rearrangements provide the mechanistic basis for tetrapolar-to-bipolar transitions but do not inherently drive shifts in breeding systems. Instead, whether a lineage remains tetrapolar or transitions to bipolarity likely depends on ecological and evolutionary pressures such as opportunities for outcrossing versus inbreeding, the frequency of sibling or clonal mating after colonization, dispersal limitations, and the long-term stability of chromosomal rearrangements that maintain linkage between the loci. Accordingly, species with linked MAT loci are often associated with animals or plants as commensals or pathogens, where linkage facilitates mating among siblings and provides an advantage during host colonization [9,28,81,138,139]. A similar scenario has been proposed for the Trichosporonales, where MAT locus fusion may have originated in a host-associated ancestor and persisted in saprobic descendants [18]. Our results, however, show that bipolarity is not restricted to pathogens: Cryptococcus sp. 3 (isolated from bark beetles) and Kwoniella fici (isolated from fig) harbor fused MAT loci with no evidence of pathogenicity to date. This suggests that restricted partner availability—whether in insect-associated microhabitats or free-living clonal populations—may favor MAT loci fusion independently of pathogenicity. Once established, bipolar systems may be difficult to revert, and their stability could in turn facilitate the persistence or emergence of pathogenic lifestyles by ensuring mating success under the constrained conditions typical of infection.

Outlook

Our characterization of MAT loci gene content and organization allowed us to infer putative breeding systems (heterothallism versus homothallism) across species. However, determining how often sexual reproduction and recombination occur in nature, and whether mating involves primarily outcrossing or inbreeding, will require population-level analyses such as linkage disequilibrium decay [140]. Unlike basidiomycete macrofungi, where sexual structures can be directly observed in the field, or smut fungi, where disease symptoms visibly reflect the production of sexual spores, most Tremellomycetes are inconspicuous environmental isolates or parasites of lichens and other fungi, complicating the isolation of compatible mating partners. Their cryptic nature has led to many species being described from a single isolate [48], hindering mating assays. Progress will require expanded taxon and population-level sampling [48,50], particularly for nonpathogenic species, to clarify reproductive strategies, transitions in mating systems, and potentially capture changes in MAT locus organization as they occur. Uncovering the hidden diversity of Cryptococcus species—particularly in their likely African origin [50]—will also be crucial for linking ecology, speciation, and adaptation with the evolutionary forces shaping MAT locus evolution and reproductive transitions, with possible implications for pathogenic emergence.

Materials and methods

Strains and media

Strains were routinely grown on YPD medium (10 g/L yeast extract, 20 g/L Bacto Peptone, 20 g/L dextrose, and 20 g/L agar) unless indicated otherwise. C. neoformans and C. deneoformans strains were incubated at 30 °C, while other Cryptococcus and Kwoniella species strains were grown at room temperature (20–23 °C). A complete list of strains is provided in S1 Appendix. Strain NCYC1536 was obtained from the National Collection of Yeast Cultures (Norwich, UK), and strain CMW60451 was isolated in October 2022 from a bark beetle (Lanurgus sp.) infesting twigs of Widdringtonia cedarbergensis in the Cederberg Mountains, South Africa.

Genomic DNA extraction

High-molecular-weight DNA was extracted with a cetyltrimethylammonium bromide extraction as previously described [108], minimizing DNA shearing during sample preparation. DNA quality was evaluated by determining the A260/A280 and A260/A230 ratios on NanoDrop spectrophotometer (Thermo). Integrity and fragment size were analyzed using clamped homogeneous electric field electrophoresis, and gDNA for short-read whole-genome sequencing (Illumina) was extracted with a phenol:chloroform-based protocol, both as previously described [49]. DNA concentration was measured with Qubit dsDNA Assay Kits (Invitrogen) on the Qubit fluorometer.

Genome sequencing

Whole-genome sequencing was performed with PacBio, Nanopore, and Illumina technologies. For K. heveanensis BCC8398 and K. mangrovensis CBS10435, Illumina sequencing at the Broad Institute Genomics Platform utilized “fragment” and “jumping” libraries, prepared and sequenced as previously described [49]. All other Illumina sequencing was conducted at the Duke University Sequencing and Genomic Technologies (DUSGT) Core, with libraries prepared with Kapa HyperPlus kit and sequenced as paired-end 2 × 150 bp reads on various Illumina platforms. For PacBio sequencing, 15–20 kb insert-size libraries were prepared and run on a PacBio RS II or Sequel (2.0 chemistry) system at the DUSGT. Nanopore sequencing was carried out in-house. Single-strain libraries were prepared with either the SQK-LSK108 or SQK-LSK110 kit, whereas up to three DNA samples were barcoded with the SQK-LSK109 and EXP-NBD103/EXP-NBD104, or SQK-NBD114.24 kits. Libraries were prepared following the manufacturer’s protocols, either individually or pooled, and sequenced on R9 (FLO-MN106) or R10 (FLO-MIN114) flow cells for 48–72 h at default voltage on a MinION Mk1B or Mk1C system. The MinION software version current at the time of each run was applied. Further details on sequencing platforms, basecalling, and demultiplexing are provided in S1 Appendix.

Genome assembly

Genomes were assembled with Canu [141] using Nanopore or PacBio sequencing data and default parameters. Assembly accuracy was improved through initial error correction with Medaka (https://github.com/nanoporetech/medaka) for Nanopore-based assemblies, followed by up to five rounds of iterative polishing with Pilon v1.22 [142] (--fix all) using Illumina reads aligned to the first pass-polished assembly with BWA-MEM v0.7.17-r1188 [143]. Pilon iterations were stopped early if no additional changes were detected, based on the absence of reported corrections in each round. Contigs containing only rDNA sequences, detected by Barrnap (https://github.com/tseemann/barrnap) (--kingdom euk), or those classified as mitochondrial DNA, were excluded from the final nuclear genome assembly. To confirm assembly completeness and evaluate telomeric regions, Nanopore/PacBio and Illumina reads were realigned to the Canu-corrected assembly using minimap2 v2.9-r720 [144] and BWA-MEM, respectively, and read coverage profiles were examined in the Integrative Genomics Viewer [145]. Draft assemblies for K. europaea PYCC6162 and K. botswanensis CBS12717 were generated with SPAdes v3.15.3 using default settings, while assemblies for K. heveanensis BCC8398 and K. mangrovensis CBS10435 were constructed with Allpaths [146]. Genome assemblies and raw sequencing data were deposited in DDBJ/EMBL/GenBank under the BioProject numbers listed in S1 Appendix, which also provides details on assembly parameters for each genome.

Gene prediction and annotation

Gene models were predicted using either BRAKER2 v2.1.5 [147] or Funannotate v1.8.9 (https://github.com/nextgenusfs/funannotate), as previously described [20,23,49]. Prior to gene prediction, assemblies were masked with RepeatMasker v4.1.4 in sensitive mode, applying a custom library that combined CryptococcusKwoniella repeat families generated with RepeatModeler2 v2.0.4 [49], centromere-associated TEs, and curated Cryptococcus elements compiled from published sources [148] and represented in Dfam. Manual inspection and correction were performed only for mating-type gene models, guided by ortholog alignments, and included adjustments to exon–intron boundaries or addition of small ORFs missed by automated pipelines (e.g., MF pheromone genes). These curated annotations were incorporated into the final GFF files and are part of the genome submissions deposited in NCBI. Functional annotation was added to the final gene models using the Funannotate “annotate” module, which integrated data for PFAM and InterPro domains, Gene Ontology terms, fungal transcription factors, Cluster of Orthologous Genes (COGs), secondary metabolites, Carbohydrate-Active Enzymes (CAZYmes), secreted proteins, proteases (MEROPS), and Benchmarking Universal Single-Copy Orthologs (BUSCO) groups. InterPro domain data, COG annotations, and secondary metabolite predictions were generated with InterProScan v5.55-88.0, eggNOGmapper v.2.1.7 (eggNOG DB version: 5.0.2) [149], and AntiSMASH v6.1.0 [150], and then passed to Funannotate annotate with the options --iprscan, --eggnog, and --antismash. Specific parameters are provided in S1 Appendix.

Ortholog identification and sequence alignment

To construct the phylogenomic data matrix, single-copy orthologs (SC-OGs) were identified across Cryptococcus and Kwoniella species and three outgroups (Tremella mesenterica ATCC28783, GCA_004117975.1; Saitozyma podzolica DSM27192, GCA_003942215.1; and Bullera alba JCM2954, GCA_001600095.1) using OrthoFinder v3.0.1b1 with options: -M msa -S diamond_ultra_sens -I 1.5 -M msa -A mafft -T fasttree -t 48 -a 6. This identified 3,086 SC-OGs shared across species. Their amino acid sequences were individually aligned with MAFFT v7.310 [151] (--localpair --maxiterate 1000) and trimmed with TrimAl v1.4.rev22 [152] (-gappyout -keepheader). The same approach was applied to construct the phylogenomic data matrix for C. neoformans strains representing different VN groups, with three C. deneoformans outgroups, yielding 5,439 shared SC-OGs, which were then aligned and trimmed as above.

Species phylogeny, estimation of topological support, and gene genealogies

The updated phylogeny of Cryptococcus and Kwoniella species (Figs 1A and S1) was inferred using two complementary strategies: (i) concatenation-based partitioned ML phylogeny reconstruction in IQ-TREE v2.1.6 [153], and (ii) gene-based coalescence analysis in ASTRAL v5.7.8 [154]. For the concatenation approach, amino acid alignments of 3,086 SC-OGs were combined into a partitioned supermatrix (52 taxa, 3,086 partitions, 1,686,691 sites) using the “-p” option in IQ-TREE. The edge-linked proportional partition model [155] was applied to account for differences in evolutionary rates across partitions, and the best-fit substitution model for each partition was identified with ModelFinder [156] under the Bayesian information criterion. The ML tree was inferred with the parameters --seed 54321 -m MFP -msub nuclear -B 1000 -alrt 1000 -T 14, incorporating 1,000 ultrafast bootstrap (UFBoot) replicates [157] and Shimodaira–Hasegawa approximate likelihood ratio tests (SH-aLRT) for branch support assessment. For the coalescence-based approach, ML gene trees were constructed for each SC-OG alignment using IQ-TREE’s “-S” option, which performs both model selection and tree inference for individual alignments. These trees were then input to ASTRAL under default settings to reconstruct the species phylogeny. Quartet support values were computed with the -t 2 option, providing quartet support metrics for the primary topology (q1) and alternative topologies (q2, q3), and local posterior probability (LPP) support. Genealogical concordance and topological support for branches in the concatenated ML tree were evaluated with gene concordance factor (gCF) and site concordance factor (sCF) metrics in IQ-TREE. This analysis utilized both the best-scoring concatenated ML tree (concat.treefile) and the set of gene trees (loci.treefile), with options -t --gcf -p --scf 100. The concatenation-based approach was also applied to resolve relationships among C. neoformans VN lineages (Fig 4A). For gene genealogies, amino acid sequences of individual genes of interest were extracted from OGs identified by OrthoFinder, manually inspected, and reannotated as needed. Alignment, trimming, and ML reconstruction were performed as described above. Trees were visualized and annotated with iTOL v7. Detailed model parameters are provided in the tree files accessible at https://doi.org/10.5281/zenodo.16987438.

Synteny analyses, MAT loci delineation, and centromere identification

Conserved synteny blocks in pairwise genome comparisons were identified with SynChro [105], using a delta parameter of three for high stringency. Comparisons in Figs 6A, 8A and S9A employed the K. shandongensis genome as reference. Detailed linear synteny plots of chromosomes and specific genomic regions, including MAT loci, were generated with EasyFig v2.2.2 [158] using BLASTN. Centromeres positions were determined in silico as previously described [49], combining detection of centromere-associated LTR elements and synteny analysis. Centromere lengths were defined as the intergenic regions between flanking centromeric genes (S1 Appendix). MAT loci were initially identified by BLAST searches, with C. neoformans-derived MAT genes and their flanking proteins as queries. The HD loci were defined as the regions spanning the HD1 and HD2 genes. The P/R loci were delineated based on structural comparisons between mating types. For species with available strains of opposite mating types, the P/R locus was defined as the region where synteny between the mating types is disrupted, with boundaries set at the points where synteny is restored. For species with only a single strain, the P/R locus length was inferred from the distance between P/R-flanking genes, as determined from comparisons involving opposite mating types.

To infer the origin of structural rearrangements observed in the P/R locus in Kwoniella (S5A Fig), we manually inspected the orientation and synteny of P/R-associated genes across mating types and species. Rearrangement breakpoints were approximated by identifying changes in gene order and orientation between alleles, supported by the presence of gene fragments (e.g., partial NDC80, STE12, BSP3) and conserved synteny in the opposite mating type. These rearrangements were reconstructed by inverting or repositioning affected gene blocks in the more derived a2 allele to match the putative ancestral configuration in the more conserved a1 allele.

Statistical analyses of MAT loci lengths across species, clades, or mating types were conducted with Python3 with Pandas, Seaborn, Matplotlib, and SciPy libraries. Differences between the two groups were assessed using the two-sided Mann–Whitney U test. Scripts and raw data are available at https://doi.org/10.5281/zenodo.16987438. Plots were refined for clarity by adjusting color schemes, labels, and adding features in Adobe Illustrator.

Analysis of gene content in MAT loci, frequency across species, and gene essentiality classification

A curated gene presence/absence matrix was constructed after delineating MAT loci across species. Genes were categorized as “present” (1), “absent” (0), “unclear” (?), or “pseudogene” (pseudo). Unclear cases indicated instances where gene presence could not be conclusively determined due to uncertain locus boundaries. For frequency calculations, pseudogenes were treated as present and unclear cases as absent. The presence of either SXI1 or SXI2 was scored as present and combined for frequency estimates. Information on gene essentiality was inferred from C. neoformans H99, based on experimentally validated studies [34,159,160] or predictions from a high-throughput transposon mutagenesis (TN-seq) study [92]. Gene essentiality in other species was extrapolated solely from H99, and genes without H99 orthologs were categorized as “unknown” due to the lack of supporting data. A custom R script (0_calculate_gene_presence_frequency.R) analyzed and plotted gene presence frequencies. The matrix was also visualized as a heatmap (1_plot_gene_matrix_based_on_frequency.R), with genes ordered by frequency. SXI1 and SXI2 are shown as separate columns in the heatmap, but their ranking reflects their combined frequency. Analyses and visualizations were conducted in R with dplyr, ggplot2, ComplexHeatmap, and circlize packages. Gene presence/absence was further analyzed with a Python script (2_compare_MAT_gene_content_across_clades.py), which calculated clade-specific and pairwise presence, identified genes shared across specified clades, and highlighted clade-specific genes. Input matrices are provided in S2 and S3 Appendices and scripts are available at https://doi.org/10.5281/zenodo.16987438.

Base composition and codon-bias analyses

Genome-wide GC content and deviations from the mean GC% were analyzed with a custom Python script (00_gc_content_analysis_and_plots.py), employing a nonoverlapping sliding window (1 kb for genome scans; 0.25 kb for zoomed-in views). Regions of interest (e.g., P/R loci) were highlighted from BED file coordinates. Plots were further refined in Adobe Illustrator for producing Fig 2. To assess whether the P/R locus exhibited lower GC content than genome-wide averages, a second script (01_gc_content_analysis.py) compared P/R locus GC% to 1,000 randomly sampled genome-wide regions of equal size (excluding the P/R locus). A one-sample t test assessed whether the P/R locus GC content differed from the sampled mean, while a permutation test (10,000 iterations) generated a null distribution to calculate P-values and Z-scores. The analysis was automated with a shell script (01a_run_script_vs_wg.sh), consolidating results into a summary table (S4 Appendix).

To assess whether GC depletion within the P/R locus reflects reduced gBGC or relaxed selection on codon usage, we performed a four-step Python3-based analysis (NumPy, pandas, SciPy, Biopython): (i) neutral noncoding GC comparison, (ii) coding third-position composition (AT3 versus GC3), (iii) codon-bias indices (Fop, Nc, ENC, ΔENC), and (iv) CAI analyses controlling for GC3 and CDS length (details in S1 Text). As preparation for CAI-related analyses, species-specific CAI reference sets were generated with a custom script (02_build_cai_reference.py) in “rp_only” mode, selecting ribosomal-protein genes as proxies for highly expressed, codon-optimized, genes [161,162], since RNA-seq data for these species was unavailable. The script counted codons and generated a CAI weight table in which, within each synonymous family, relative adaptiveness weights (w) were normalized so that the most frequent codon had w = 1 and others 0 < w ≤ 1. Per-gene CAI values were then computed with 03_compute_cai_per_gene.py, which reads each species’ CAI weight table and CDS annotations and returns gene-level CAI (geometric mean of w across codons; start/stop excluded; nonACGT ignored; Met/Trp treated as neutral set to w = 1).

To test for GC depletion in the P/R region, noncoding GC within P/R (intergenic plus introns) was compared to same-chromosome, length-matched nulls with script 04_neutral_non_coding_GC.py using 10,000 P/R-excluded windows (--n_null 10000 --no_overlap --min_gap 1000). The script reports ΔGC (gc_intergenicnull_mean_intergenic), a standardized z, and a one-sided empirical permutation p-value (PR ≤ null).

Coding third-position composition (AT3 versus GC3) was analyzed with script 05_coding_composition_and_codon_bias.py by counting AT- versus GC-ending third positions inside P/R and background regions (same chromosome, outside P/R) and testing the 2 × 2 table by χ2 with Fisher’s exact fallback. Using the same script and the CAI weights above, codon-bias indices were calculate: Fop (frequency of optimal codons based on the CAI reference); Nc (effective number of codons, representing the observed codon usage bias in a gene); ENC (the expected Nc based solely on the gene’s GC3, calculated from the GC3–Nc relationship described by Wright (1990) [163]); and ΔENC = Nc − ENC (representing codon bias beyond that expected from GC3 composition alone). To improve the stability of Nc estimates, particularly for short CDS, we applied a small pseudocount (0.5 per codon), excluded very small synonymous families, bounded intermediate calculations to avoid division by near-zero values, and constrained Nc values to the theoretical range of 20 (maximum bias) to 61 (no bias) before computing ΔENC. P/R versus background contrasts were evaluated as median differences using Mann–Whitney tests with 10,000 permutations.

Finally, CAI differences independent of nucleotide composition were tested using three complementary analyses implemented in script 06_cai_vs_gc3_controls.py (detailed in S1 Text), using the following parameters: --n_perm 10000, --gc3_tol 0.03, --len_tol 0.10, and --k_match 5. Briefly, we (i) fit CAI ~ inside_pr +  Δ(exon-intron GC) + log(length_cds) to all genes using ordinary-least-squares (OLS) with heteroskedasticity-consistent (HC1) robust standard errors; inside_pr = 1 for CDS fully within the P/R window and 0 for same-chromosome genes outside (boundary overlaps excluded); (ii) we trained CAI ~ Δ(exon-intron GC) + log(length_cds) on background genes only, computed residuals for all genes, and tested median(PR residual) − median(BG residual) with a one-sided permutation test (10,000 shuffles; alternative PR < BG); and (iii) we matched each P/R locus gene to same-chromosome controls within ±0.03 GC3 and ±10% CDS length (dropping unmatched P/R genes), computed CAI_match_median = median[CAI_PR − mean(CAI_BG matches)], and used a sign-flip permutation test (10,000 flips; alternative PR < BG). Fop was analyzed identically to step (iii) with the same matching tolerances and permutation settings. All scripts are available at https://doi.org/10.5281/zenodo.16987438, and per-species summaries and statistical tests are given in S4 Appendix.

Variant analysis and SNP distribution in C. neoformans MATa strains

Multi-genome variant analysis was performed with the Snippy pipeline v4.6.0 (https://github.com/tseemann/snippy) using the newly assembled genome of C. neoformans Ftc555-1 (VNBI) as the reference, with parameters: --cpus 10, --unmapped, --mincov 10, and --minfrac 0.9. Paired-end reads were obtained from the NCBI SRA database (S1 Appendix). Three datasets were independently used: (i) the entire genome, (ii) only the MAT locus region (chr_5:153,773-277,230), and (iii) only chr. 5. The MAT locus and chr. 5 were extracted from the reference genome to generate region-specific GenBank files for Snippy to identify SNPs and construct region-specific phylogenies. Visualization of SNV distributions was based on the whole-genome dataset and included both chromosome-wide and zoomed-in views of specific regions. A custom Python script processed merged SNV datasets, computed total SNP counts for each strain, and generated gradient color-coded SNP density plots. Final composite figures were refined using Adobe Illustrator. Core SNP alignments (core.aln) from each dataset were used for phylogenetic reconstruction with IQ-TREE2, incorporating 10,000 UFBoot replicates and SH-aLRT tests for branch support. Scripts for SNV analysis and visualization are available at https://doi.org/10.5281/zenodo.16987438.

Chromosome composition analysis of K. mangrovensis progeny

To characterize chromosome composition in K. mangrovensis progeny, a combined nuclear reference genome was constructed from assemblies of the two parental strains, CBS8507 and CBS10435. For CBS8507, the original assembly consisted of four contigs, two of which represented the same chromosome fragmented at the rDNA array. To streamline downstream analyses and facilitate interpretation, the two contigs were merged into a single scaffold: orientation was first established from rDNA gene direction; the corresponding ends were aligned and trimmed; and a 5-N gap was inserted to join them, yielding a finalized three-chromosome assembly. For CBS10435, the draft assembly of 37 contigs was reordered and reoriented to match the CBS8507 3-chromosome assembly using D-GENIES [164], which employs minimap2 for genome alignment. Raw Illumina paired-end reads from selected progeny were then mapped to a combined reference with the sppIDer pipeline [165], which sequentially mapped reads, applied stringent quality filtering (MQ > 3), and generated depth-of-coverage plots. Chromosome number and ploidy were estimated by integrating spider coverage data with flow cytometry results.

Mating assays, phenotyping on mating-inducing media, and microscopy

Mating assays with K. mangrovensis strains CBS8507 and CBS10435 were performed as previously described [66,67]. Equal amounts of cells from each strain were mixed on V8 agar (10 g/L yeast extract, 20 g/L Bacto Peptone, 20 g/L dextrose, 20 g/L agar; pH 5) or corn meal agar (CMA; 15 g/L corn meal agar, 5 g/L agar) and incubated in the dark at room temperature (20–23 °C) for up to 1 month. Plates were regularly monitored for the development of mating structures. Random F1 progeny were recovered by microdissecting individual cells from cell clusters [79] embedded in or on the surface of the mating media (V8 and CMA) at the edges of mating patches. A total of 51 progeny were collected, grown axenically on YPD plates, and stored as glycerol stocks at −80 °C. Selected F1 progeny (MP11, MP15, MP16, MP18, MP49, MP50, MP51, MP54, and MP59) were assessed for hyphal formation and mating structures on V8 pH5 and CMA, compared to the parental cross (CBS8507 × CBS10432) and the solo culture of each parental strain. Potentially compatible strain pairs from Cryptococcus species lacking documented sexual reproduction (IND107 × CBS11718; CBS11687 × 7685027; and DSM108351 × NCYC1536) were tested for mating ability under conditions known to induce mating in C. neoformans and C. deneoformans (V8 media, dark, room temperature) for up to a month, with regular monitoring [79]. Solo cultures of each strain were also evaluated for their ability to reproduce sexually without a compatible partner. Similar assays were performed for strains of newly described species lacking known compatible partners. All mating assays were repeated three separate times to ensure consistent phenotypic observations. To assess hyphal growth, basidia, and spores, the edges of yeast colonies and mating patches were examined and photographed with a Zeiss Axio Scope.A1 microscope equipped with an Axiocam Color camera and ZEN Lite v3.4 software.

Analysis of K. mangrovensis progeny by PCR-restriction fragment length polymorphism (RFLP) analysis

Kwoniella mangrovensis has only three chromosomes. To analyze the F1 progeny from the CBS8507 × CBS10432 crosses, six primer pairs (one per chromosomal arm) were designed for the initial screening. Primers were manually designed to produce distinct parental patterns after PCR product digestion with specific restriction enzymes. As a heterozygosity control, PCR products from a 1:1 DNA mixture of both parental strains were amplified for each primer pair, followed by digestion with the corresponding restriction enzyme. Results and primer sequences are provided in S5 Appendix.

Fluorescence-activated cell-sorting (FACS)

FACS was performed as previously described [166] to determine ploidy of K. mangrovensis CBS8507, CBS10432, and nine selected progeny. Strains were grown overnight at room temperature (21–23 °C) on YPD medium, harvested, and washed with PBS. Cells were fixed in 2 ml of 70% ethanol at 4 °C overnight, washed with 1 ml of NS buffer (10 mM Tris-HCl pH 7.2, 250 mM sucrose, 1 mM EDTA pH 8.0, 1 mM MgCl2, 0.1mM CaCl2, 0.1 mM ZnCl2, 0.4 mM phenylmethylsulfonyl fluoride, and 7 mM β-mercaptoethanol), and then stained with 5 μl of propidium iodide (0.5 mg/ml) in 180 μl NS buffer with 20 μl of RNase (10 mg/ml) at 4 °C overnight. Lastly, 50 μl of stained cells were diluted in 2 ml of 50 mM Tris-HCl pH 8.0 and sonicated for 1 min before analysis at the Duke Cancer Institute Flow Cytometry Shared Resource. Data were collected from 10,000 cells using the FL1 channel on a Becton-Dickinson FACScan and analyzed with FlowJo software. Strains JEC21 and XL143 served as haploid and diploid controls, respectively [167].

Supporting information

S1 Fig. Cryptococcus and Kwoniella species phylogenies.

Both trees were inferred using a dataset of 3,086 single-copy genes shared across all species and three outgroups (depicted in gray). (A) Phylogeny inferred using a concatenation-based approach, corresponding to the tree in Fig 1A, but with all branch support and concordance values displayed (SH-aLRT, UFBoot, gCF, and sCF). (B) Tree topology inferred using a coalescence-based approach (ASTRAL), with local posterior probability (LPP) values shown for each branch. Quartet values for the main topology (q1) and alternative topologies (q2 and q3) are included for branches where q1 < 0.4, reflecting some uncertainty considering the expected value of 0.33 for a hard polytomy.

https://doi.org/10.1371/journal.pbio.3003417.s001

(PDF)

S2 Fig. Assessing sexual reproduction in Cryptococcus species.

All micrographs are from cultures incubated for 2 weeks in the dark at room temperature on the indicated media. For each panel, the left image shows the colony edge, and the right image shows a higher-magnification view of cell or hyphal morphology from the same culture. (A) C. decagattii: solo cultures of strains 7685027 (MATα) and CBS11687 (MATa) produced only yeast cells (right panels). In contrast, the α × a cross generated hyphae at the colony margin, with basidia and a few elongated basidiospores (arrowheads). Long spore chains were not observed, possibly due to rapid spore discharge. (B) C. tetragattii: solo cultures of strains IND107 (MATα) and CBS11718 (MATa) produced only yeast cells (right panels), and the α × a cross did not form hyphae or sexual structures, even after prolonged incubation (>1 month). (C) Solo cultures of Cryptococcus sp. 3 (CMW60451) and C. depauperatus (CBS7841) on YPD and V8 pH 5 media. C. depauperatus produced abundant hyphae on both media; however, basidia (arrowhead) and spore chains (brackets) were only observed under V8 conditions. In contrast, Cryptococcus sp. 3 grew primarily as yeast on YPD, occasionally forming pseudohyphae on V8 pH 5, but no sexual structures were detected even after extended incubation (>1 month). (D) Cryptococcus sp. 4: solo cultures of NCYC1536 (a1b1) and DSM108351 (a2b2) grew as yeast (right panels). The a1b1 × a2b2 cross produced limited hyphae (right panel, zoomed view), but these were incipient and nonrecurrent, and no sexual structures were observed even after >2 months. Scale bars: 200 μm (colony edge images); 10 μm (right panels), except for panel D bottom right = 25 μm.

https://doi.org/10.1371/journal.pbio.3003417.s002

(TIF)

S3 Fig. Structure and genomic context of the HD mating-type locus in tetrapolar Cryptococcus and Kwoniella species.

This supplementary figure spans five pages and contains a total of 15 panels (labeled A–O), with three consecutive panels per page. For each group of three panels, the top-left panel shows a synteny view of the full chromosomes, highlighting the chromosomal location of the HD locus; the bottom panel provides a zoomed-in synteny view (~200 kb) centered on the HD locus; and the top-right panel displays a simplified phylogenetic tree, providing contextual information for the species included in the synteny analyses. For cross-referencing, HD-associated genes in the zoomed-in panel are colored gold if the corresponding ortholog in Cryptococcus pathogens is located within the MAT locus or shown in a darker shade when positioned in the immediate flanking regions. The P/R allele of each strain (a1 or a2) is indicated on the left. Chromosomes inverted relative to their original assembly orientations are marked with asterisks. In panel G, only one representative of K. europaea and K. botswanensis is shown, and in panel M, only one representative of K. heveanensis is included, as the genomes of their mating-type counterparts are not assembled at the chromosome level; the scaffolds containing the HD locus are, however, included in the zoomed-in view. Other features are annotated as shown in the key.

https://doi.org/10.1371/journal.pbio.3003417.s003

(PDF)

S4 Fig. Structure and genomic context of the P/R mating-type locus in tetrapolar Cryptococcus and Kwoniella species.

This supplementary figure spans five pages and contains a total of 15 panels (labeled A–O), with three consecutive panels per page. For each group of three panels, the top-left panel shows a synteny view of the full chromosomes, highlighting the chromosomal location of the P/R locus; the bottom panel provides a zoomed-in synteny view (~200 kb) that includes the full P/R locus; and the top-right panel displays a simplified phylogenetic tree, providing contextual information for the species included in the synteny analyses. For cross-referencing, P/R-associated genes in the zoomed-in panel are colored teal if the corresponding ortholog in Cryptococcus pathogens is located within the MAT locus, shown in a darker shade when positioned in the flanking regions, or colored bright green if the gene is found within the P/R locus of some species. The P/R allele of each strain (a1 or a2) is indicated on the left. Chromosomes inverted relative to their original assembly orientations are marked with asterisks. In panels A and D, the GC content is depicted as the deviation from the genome average, calculated in 0.5 kb nonoverlapping windows. In panel G, only one representative of K. europaea and K. botswanensis is shown, and in panel M, only one representative of K. heveanensis is included, as the genomes of their mating-type counterparts are not assembled at the chromosome level; the scaffolds containing the P/R locus are, however, included in the zoomed-in view. Other features are annotated as shown in the key.

https://doi.org/10.1371/journal.pbio.3003417.s004

(PDF)

S5 Fig. Predicted rearrangements leading to the extant P/R configuration in Kwoniella species within clades E/F and clade G.

(A) The extant P/R locus configuration in K. mangrovensis is inferred to derive from a putative ancestral state through three inversion events: (INV1) involving STE3 and STE12, truncating STE12 in the a2 allele; (INV2) relocating DBP10 and REI1 from the locus edge to its center; and (INV3) moving NDC80 from the edge to the middle of the locus. (B) The extant P/R locus configuration in K. heveanensis, inferred from a putative ancestral state involving four inversion events: (INV1) involving STE3 and STE12, truncating STE12 in the a2 allele; (INV1b) inverting only the pheromone gene; (INV2) relocating DBP10 and REI1 from the locus edge to its center; (INV3) relocating BSP3-NDC80 to the middle of the locus; and (INV4) relocating UBP6 to the center of the locus. INV1 and INV2 are likely ancestral in Kwoniella. (C) Gene genealogies reveal trans-specific polymorphism for ATG7, UBP6, and BSP3 within clade G species only. Genes that required manual correction are marked with an asterisk.

https://doi.org/10.1371/journal.pbio.3003417.s005

(PDF)

S6 Fig. MAT locus structure in Cryptococcus deneoformans.

Synteny analysis is shown for five strains organized by mating type. Strain JEC20a is an F1 progeny of the cross between isolates NIH12α and NIH433a, while JEC21α is its congenic partner, generated through 10 rounds of backcrossing to JEC20a. Therefore, the MATα allele in JEC21α was inherited from NIH12α, and the MATa allele in JEC20a was inherited from NIH433a. The genes BSP3 and IKS1 are no longer part of the MAT locus in this species. Chromosomes inverted relative to their original assembly orientations are marked with asterisks.

https://doi.org/10.1371/journal.pbio.3003417.s006

(PDF)

S7 Fig. MAT locus structure in Cryptococcus gattii species complex.

This supplementary figure spans 2 pages. (A) Synteny analysis with strains organized by mating type and, within each mating type, sorted according to their phylogenetic relationships. The analysis shows that the MATa configuration is highly variable across species, with each species exhibiting a distinct organization. In contrast, the MATα structure is more conserved across species. (B) Synteny analysis with strains grouped by species to compare differences between mating types and define the exact boundaries of the MAT locus based on synteny conservation. Note that no MATa strain of the VGV lineage has been isolated to date. In both panels, chromosomes inverted relative to their original assembly orientations are marked with asterisks.

https://doi.org/10.1371/journal.pbio.3003417.s007

(PDF)

S8 Fig. Phylogenies of BSP3, PRT1, SPO14, and CAP1 within the predicted MAT locus of Cryptococcus sp.3.

In pathogenic Cryptococcus species, these genes exhibit trans-specific polymorphism, characterized by mating-type-specific clustering, with a and α alleles forming distinct clades. An exception is observed in BSP3 from C. deneoformans (marked with asterisks), where mating-type specificity has been lost, representing a derived state. In contrast, protein sequences from Cryptococcus sp. 3 (highlighted in bold) do not group with either the a or α allele-specific clusters and instead occupy distinct positions outside these groups. Maximum likelihood phylogenies were constructed using protein sequences. For clarity, sequences from other Cryptococcus and Kwoniella species are collapsed, and well-supported nodes (≥90% SH-aLRT/UFBoot support) are indicated by filled circles. Scale bars represent the number of substitutions per site.

https://doi.org/10.1371/journal.pbio.3003417.s008

(PDF)

S9 Fig. Chromosome number reduction in Cryptococcus sp. 3 (CWM60451) involves centromere inactivation via loss of LTR-rich regions.

(A) The karyotype of K. shandongensis (with 14 chrs.) served as the reference for reconstructing synteny blocks in pairwise comparisons with representative Cryptococcus species of clades A (C. neoformans H99), B (C. amylolentus CBS6039), C (Cryptococcus sp. CMW60451), and D (Cryptococcus sp. DSM108351). Each K. shandongensis chromosome is assigned a distinct color, which defines the color scheme for homologous synteny blocks in the corresponding chromosomes of other species. Centromeres are labeled “CEN” and colored according to their associated reference chromosome. Centromeres inferred to result from intercentromeric recombination are shown with two colors, reflecting ancestry from two distinct K. shandongensis centromeres. Inactivated centromeres (“iCEN” or “ic”) are marked with red arrowheads. (B) Linear chromosome plots show synteny conservation among these species and detail the region of centromere inactivation (iCEN) in Cryptococcus sp. 3. Centromeres are marked with black boxes with white circles and repeat-rich regions are highlighted in pink.

https://doi.org/10.1371/journal.pbio.3003417.s009

(PDF)

S10 Fig. Nuclear genome composition of Kwoniella mangrovensis progeny reveals substantial aneuploidy and recombination.

Sequencing read-depth coverage and inheritance patterns of progeny derived from a CBS8507 (a1b1) × CBS10435 (a2b2) sexual cross. For each progeny, sequencing coverage plots (normalized to the genome-wide average coverage) are color-coded according to each parent’s contribution, as shown in the key. Haplotypes blocks inferred from SNP data are overlaid for each chromosome for comparison, reflecting instances of recombination or loss of heterozygosity (LOH). In some strains (MP18, MP50, MP59, and MP54), coverage analysis suggests potential changes in ploidy in a subset of the sequenced cell population. This is indicated by skewed read proportions favoring one parent genome over the other while still retaining the corresponding haplotype, suggestive of genomic instability.

https://doi.org/10.1371/journal.pbio.3003417.s010

(PDF)

S11 Fig. FACS analysis of Kwoniella mangrovensis parental strains CBS8507 and CBS10435, and their progeny.

The C. neoformans strain JEC21 and XL143 served as haploid (1n), and diploid (2n) controls, respectively.

https://doi.org/10.1371/journal.pbio.3003417.s011

(PDF)

S12 Fig. Self-filamentation phenotype of Kwoniella mangrovensis progeny.

Parental strains CBS8507 and CBS10435 (grown individually and in co-culture) and their recovered progeny were cultivated on V8, pH 5.0 at room temperature. Self-filamentation was assessed after 2 weeks of incubation. While neither parental strain exhibited self-filamentation in solo culture, their co-culture produced hyphal filaments and basidia.

https://doi.org/10.1371/journal.pbio.3003417.s012

(TIF)

S1 Text. GC depletion at the P/R locus in tetrapolar Cryptococcus is best explained by reduced GC-biased gene conversion.

https://doi.org/10.1371/journal.pbio.3003417.s013

(PDF)

S2 Text. Karyotype reduction in Cryptococcus sp. 3 is independent of P/R–HD linkage.

https://doi.org/10.1371/journal.pbio.3003417.s014

(PDF)

S1 Appendix.

Genome assembly, genomic features, and information on raw sequencing data generated in this study. (A) List of Cryptococcus and Kwoniella isolates used in this study and summary of genome assembly statistics and other genomic features. (B) Genome sequencing, assembly, and polishing approaches. (C) NCBI accession numbers of each genome and raw read data generated and used in this study. (D) SRA accession numbers of Illumina read data of C. neoformans MATa strains used for variant calling and SNP distribution. (E) Contig length, centromere coordinates, and telomeric sequences of genome assemblies generated and analyzed in this study.

https://doi.org/10.1371/journal.pbio.3003417.s015

(XLSX)

S2 Appendix.

Mating-type (MAT) loci regions and genes within MAT in Cryptococcus and Kwoniella. (A) Summary of the MAT loci regions in bipolar and tetrapolar Cryptococcus and Kwoniella species. (B) List of genes within the P/R locus in tetrapolar Cryptococcus and Kwoniella species and within MAT in bipolar species and their essentiality classification. (C) P/R locus size comparisons between tetrapolar Cryptococcus and Kwoniella species. (D) P/R locus size comparisons between tetrapolar Cryptococcus species. (E) MAT locus size comparisons between MATα and MATa strains of C. neoformans. (F) MAT locus size comparisons between MATα and MATa strains of the C. gattii species complex.

https://doi.org/10.1371/journal.pbio.3003417.s016

(XLSX)

S3 Appendix.

Shared and unique mating-type (MAT) genes across Cryptococcus and Kwoniella clades. (A) List of genes within MAT across Cryptococcus and Kwoniella species and clades. (B) Shared and unique genes within P/R locus in Cryptococcus and Kwoniella clades. (C) Shared and unique genes within MAT loci (i.e., including P/R- and HD-associated genes) in Cryptococcus and Kwoniella clades.

https://doi.org/10.1371/journal.pbio.3003417.s017

(XLSX)

S4 Appendix.

GC content and codon composition/usage analysis. (A) Genome-wide GC content and deviation from genome mean in clade B and D strains included in Fig 2. (B) GC content reduction in the P/R locus relative to genome-wide background (clades C and D). (C) Neutral comparison of noncoding GC in P/R locus vs. chromosome-matched background windows (Test 1). (D) Third-position codon usage (AT3 vs. GC3) in P/R locus vs. background (Test 2). (E) Codon usage and bias metrics in P/R vs. matched background genes (Tests 2 and 3). (F) Statistical tests for codon usage bias in P/R vs. background (Tests 3 and 4). (G) Tests for reduced codon adaptation (CAI) after controlling for GC3 and gene length (Test 4).

https://doi.org/10.1371/journal.pbio.3003417.s018

(XLSX)

S5 Appendix.

List of primers and PCR-RFLP analysis of K. mangrovensis sexual progeny. (A) List of primers and conditions used to perform PCR-RFLP. (B) List of F1 progeny recovered from CBS8507 × CBS10435 cross and selected for Illumina sequencing after PCR-RFLP analysis. (C) List of all F1 progeny recovered from CBS8507 × CBS10435 cross. (D) PCR-RFLP results for amplicon MP293-MP294 digested with EcoRV. (E) PCR-RFLP results for amplicon MP303-MP304 digested with EcoRV. (F) PCR-RFLP results for amplicon MP307-MP308 digested with EcoRV. (G) PCR-RFLP results for amplicon MP309-MP310 digested with EcoRV. (H) PCR-RFLP results for amplicon MP311-MP312 digested with EcoRI. (I) PCR-RFLP results for amplicon MP313-MP314 digested with ClaI.

https://doi.org/10.1371/journal.pbio.3003417.s019

(XLSX)

Acknowledgments

We thank Vikas Yadav for assistance with Nanopore data acquisition, Terrance Shea and Christina Cuomo for providing the genomes assemblies of K. mangrovensis CBS8886 and CBS10435, as well as K. heveanensis BCC8398, and the Broad Institute Genomics Platform for generating Illumina sequencing data for these three strains. We also thank Fred Dietrich for computational resources, Prof Francois Roets for providing permits to collect specimens in South Africa, and Minou Nowrousian for critical reading of this manuscript.

References

  1. 1. Otto SP, Lenormand T. Resolving the paradox of sex and recombination. Nat Rev Genet. 2002;3(4):252–61. pmid:11967550
  2. 2. de Visser JAGM, Elena SF. The evolution of sex: empirical insights into the roles of epistasis and drift. Nat Rev Genet. 2007;8(2):139–49. pmid:17230200
  3. 3. Goodenough U, Heitman J. Origins of eukaryotic sexual reproduction. Cold Spring Harb Perspect Biol. 2014;6(3):a016154. pmid:24591519
  4. 4. Hawksworth DL, Lücking R. Fungal diversity revisited: 2.2 to 3.8 million species. Microbiol Spectr. 2017;5(4):10.1128/microbiolspec.funk-0052–2016. pmid:28752818
  5. 5. Heitman J, Howlett BJ, Crous PW, Stukenbrock EH, James TY, Gow NAR. The fungal kingdom. American Society of Microbiology; 2017.
  6. 6. Begerow D, Kemler M, Feige A, Yurkov A. Parasitism in yeasts. Yeasts in natural ecosystems: ecology. Springer International Publishing; 2017. p. 179–210. doi: https://doi.org/10.1007/978-3-319-61575-2_7
  7. 7. Lee SC, Ni M, Li W, Shertz C, Heitman J. The evolution of sex: a perspective from the fungal kingdom. Microbiol Mol Biol Rev. 2010;74(2):298–340. doi: https://doi.org/10.1128/MMBR.00005-10 pmid:20508251
  8. 8. Lin X, Heitman J. Mechanisms of homothallism in fungi and transitions between heterothallism and homothallism. In: Heitman J, Kronstad JW, Taylor JW, Casselton LA, editors. Sex in fungi: molecular determination and evolutionary implications. Washington D.C.: ASM Press. 2007. p. 35–57.
  9. 9. Coelho MA, Bakkeren G, Sun S, Hood ME, Giraud T. Fungal sex: the basidiomycota. Microbiol Spectr. 2017;5(3):10.1128/microbiolspec.funk-0046–2016. pmid:28597825
  10. 10. Kües U, James TY, Heitman J. Mating type in basidiomycetes: unipolar, bipolar, and tetrapolar patterns of sexuality. In: Pöggeler S, Wöstemeyer J, editors. Evolution of fungi and fungal-like organisms. Heidelberg, Germany: Springer Berlin Heidelberg; 2011. p. 97–160.
  11. 11. James TY, Srivilai P, Kües U, Vilgalys R. Evolution of the bipolar mating system of the mushroom Coprinellus disseminatus from its tetrapolar ancestors involves loss of mating-type-specific pheromone receptor function. Genetics. 2006;172(3):1877–91. pmid:16461425
  12. 12. Yi R, Tachikawa T, Ishikawa M, Mukaiyama H, Bao D, Aimi T. Genomic structure of the A mating-type locus in a bipolar basidiomycete, Pholiota nameko. Mycol Res. 2009;113(Pt 2):240–8. pmid:19049868
  13. 13. James TY, Lee M, van Diepen LTA. A single mating-type locus composed of homeodomain genes promotes nuclear migration and heterokaryosis in the white-rot fungus Phanerochaete chrysosporium. Eukaryot Cell. 2011;10(2):249–61. pmid:21131435
  14. 14. Olson Å, Aerts A, Asiegbu F, Belbahri L, Bouzid O, Broberg A, et al. Insight into trade-off between wood decay and parasitism from the genome of a fungal forest pathogen. New Phytol. 2012;194(4):1001–13. pmid:22463738
  15. 15. Bakkeren G, Kronstad JW. Linkage of mating-type loci distinguishes bipolar from tetrapolar mating in basidiomycetous smut fungi. Proc Natl Acad Sci U S A. 1994;91(15):7085–9. pmid:7913746
  16. 16. Lengeler KB, Fox DS, Fraser JA, Allen A, Forrester K, Dietrich FS, et al. Mating-type locus of Cryptococcus neoformans: a step in the evolution of sex chromosomes. Eukaryot Cell. 2002;1(5):704–18. pmid:12455690
  17. 17. Badouin H, Hood ME, Gouzy J, Aguileta G, Siguenza S, Perlin MH, et al. Chaos of rearrangements in the mating-type chromosomes of the anther-smut fungus Microbotryum lychnidis-dioicae. Genetics. 2015;200(4):1275–84. pmid:26044594
  18. 18. Sun S, Coelho MA, Heitman J, Nowrousian M. Convergent evolution of linked mating-type loci in basidiomycete fungi. PLoS Genet. 2019;15(9):e1008365. pmid:31490920
  19. 19. Gioti A, Nystedt B, Li W, Xu J, Andersson A, Averette AF, et al. Genomic insights into the atopic eczema-associated skin commensal yeast Malassezia sympodialis. mBio. 2013;4(1):e00572-12. pmid:23341551
  20. 20. Coelho MA, Ianiri G, David-Palma M, Theelen B, Goyal R, Narayanan A, et al. Frequent transitions in mating-type locus chromosomal organization in Malassezia and early steps in sexual reproduction. Proc Natl Acad Sci U S A. 2023;120(32):e2305094120. pmid:37523560
  21. 21. David-Palma M, Sampaio JP, Gonçalves P. Genetic dissection of sexual reproduction in a primary homothallic basidiomycete. PLoS Genet. 2016;12(6):e1006110. pmid:27327578
  22. 22. Cabrita A, David-Palma M, Brito PH, Heitman J, Coelho MA, Gonçalves P. Multiple pathways to homothallism in closely related yeast lineages in the Basidiomycota. mBio. 2021;12(1):e03130-20. pmid:33593979
  23. 23. Passer AR, Clancey SA, Shea T, David-Palma M, Averette AF, Boekhout T, et al. Obligate sexual reproduction of a homothallic fungus closely related to the Cryptococcus pathogenic species complex. Elife. 2022;11:e79114. pmid:35713948
  24. 24. Wang Y-W, McKeon MC, Elmore H, Hess J, Golan J, Gage H, et al. Invasive Californian death caps develop mushrooms unisexually and bisexually. Nat Commun. 2023;14(1):6560. pmid:37875491
  25. 25. Lin X, Hull CM, Heitman J. Sexual reproduction between partners of the same mating type in Cryptococcus neoformans. Nature. 2005;434(7036):1017–21. pmid:15846346
  26. 26. Maia TM, Lopes ST, Almeida JMGCF, Rosa LH, Sampaio JP, Gonçalves P, et al. Evolution of mating systems in basidiomycetes and the genetic architecture underlying mating-type determination in the yeast Leucosporidium scottii. Genetics. 2015;201(1):75–89. pmid:26178967
  27. 27. Branco S, Carpentier F, Rodríguez de la Vega RC, Badouin H, Snirc A, Le Prieur S, et al. Multiple convergent supergene evolution events in mating-type chromosomes. Nat Commun. 2018;9(1):2000. pmid:29784936
  28. 28. Nieuwenhuis BPS, Billiard S, Vuilleumier S, Petit E, Hood ME, Giraud T. Evolution of uni- and bifactorial sexual compatibility systems in fungi. Heredity (Edinb). 2013;111(6):445–55. pmid:23838688
  29. 29. James TY. Why mushrooms have evolved to be so promiscuous: insights from evolutionary and ecological patterns. Fungal Biol Rev. 2015;29(3–4):167–78.
  30. 30. Taniguti LM, Schaker PDC, Benevenuto J, Peters LP, Carvalho G, Palhares A, et al. Complete genome sequence of Sporisorium scitamineum and biotrophic interaction transcriptome with sugarcane. PLoS One. 2015;10(6):e0129318. pmid:26065709
  31. 31. Depotter JRL, Ökmen B, Ebert MK, Beckers J, Kruse J, Thines M, et al. High nucleotide substitution rates associated with retrotransposon proliferation drive dynamic secretome evolution in smut pathogens. Microbiol Spectr. 2022;10(5):e0034922. pmid:35972267
  32. 32. Carpentier F, Rodríguez de la Vega RC, Branco S, Snirc A, Coelho MA, Hood ME, et al. Convergent recombination cessation between mating-type genes and centromeres in selfing anther-smut fungi. Genome Res. 2019;29(6):944–53. pmid:31043437
  33. 33. Duhamel M, Carpentier F, Begerow D, Hood ME, Rodríguez de la Vega RC, Giraud T. Onset and stepwise extensions of recombination suppression are common in mating-type chromosomes of Microbotryum anther-smut fungi. J Evol Biol. 2022;35(12):1619–34. pmid:35271741
  34. 34. Fraser JA, Diezmann S, Subaran RL, Allen A, Lengeler KB, Dietrich FS, et al. Convergent evolution of chromosomal sex-determining regions in the animal and fungal kingdoms. PLoS Biol. 2004;2(12):e384. pmid:15538538
  35. 35. Sun S, Yadav V, Billmyre RB, Cuomo CA, Nowrousian M, Wang L, et al. Fungal genome and mating system transitions facilitated by chromosomal translocations involving intercentromeric recombination. PLoS Biol. 2017;15(8):e2002527. pmid:28800596
  36. 36. Nichols CB, Fraser JA, Heitman J. PAK kinases Ste20 and Pak1 govern cell polarity at different stages of mating in Cryptococcus neoformans. Mol Biol Cell. 2004;15(10):4476–89. pmid:15282344
  37. 37. Sun S, Coelho MA, David-Palma M, Priest SJ, Heitman J. The evolution of sexual reproduction and the mating-type locus: links to pathogenesis of Cryptococcus human pathogenic fungi. Annu Rev Genet. 2019;53:417–44. pmid:31537103
  38. 38. Branco S, Badouin H, Rodríguez de la Vega RC, Gouzy J, Carpentier F, Aguileta G, et al. Evolutionary strata on young mating-type chromosomes despite the lack of sexual antagonism. Proc Natl Acad Sci U S A. 2017;114(27):7067–72. pmid:28630332
  39. 39. Hartmann FE, Duhamel M, Carpentier F, Hood ME, Foulongne-Oriol M, Silar P, et al. Recombination suppression and evolutionary strata around mating-type loci in fungi: documenting patterns and understanding evolutionary and mechanistic causes. New Phytol. 2021;229(5):2470–91. pmid:33113229
  40. 40. Hagen F, Khayhan K, Theelen B, Kolecka A, Polacheck I, Sionov E, et al. Recognition of seven species in the Cryptococcus gattii/Cryptococcus neoformans species complex. Fungal Genet Biol. 2015;78:16–48. pmid:25721988
  41. 41. Farrer RA, Chang M, Davis MJ, van Dorp L, Yang D-H, Shea T, et al. A new lineage of Cryptococcus gattii (VGV) discovered in the Central Zambezian Miombo Woodlands. mBio. 2019;10(6):e02306-19. pmid:31719178
  42. 42. Fisher MC, Denning DW. The WHO fungal priority pathogens list as a game-changer. Nat Rev Microbiol. 2023;21(4):211–2. pmid:36747091
  43. 43. Casalini G, Giacomelli A, Antinori S. The WHO fungal priority pathogens list: a crucial reappraisal to review the prioritisation. Lancet Microbe. 2024;5(7):717–24. pmid:38608682
  44. 44. Ginns J, Bernicchia A. Filobasidiella lutea: parasitism of Hypochnicium vellereum. Karstenia. 2000;40:49–51.
  45. 45. Findley K, Rodriguez-Carres M, Metin B, Kroiss J, Fonseca A, Vilgalys R, et al. Phylogeny and phenotypic characterization of pathogenic Cryptococcus species and closely related saprobic taxa in the Tremellales. Eukaryot Cell. 2009;8(3):353–61. pmid:19151324
  46. 46. Liu X-Z, Wang Q-M, Göker M, Groenewald M, Kachalkin AV, Lumbsch HT, et al. Towards an integrated phylogenetic classification of the Tremellomycetes. Stud Mycol. 2015;81:85–147. pmid:26955199
  47. 47. Passer AR, Coelho MA, Billmyre RB, Nowrousian M, Mittelbach M, Yurkov AM, et al. Genetic and genomic analyses reveal boundaries between species closely related to Cryptococcus pathogens. mBio. 2019;10(3):e00764-19. pmid:31186317
  48. 48. Kachalkin AV, Turchetti B, Inácio J, Carvalho C, Mašínová T, Pontes A, et al. Rare and undersampled dimorphic basidiomycetes. Mycol Progress. 2019;18(7):945–71.
  49. 49. Coelho MA, David-Palma M, Shea T, Bowers K, McGinley-Smith S, Mohammad AW, et al. Comparative genomics of the closely related fungal genera Cryptococcus and Kwoniella reveals karyotype dynamics and suggests evolutionary mechanisms of pathogenesis. PLoS Biol. 2024;22(6):e3002682. pmid:38843310
  50. 50. Coelho MA, David-Palma M, Aylward J, Pham NQ, Visagie CM, Fuchs T, et al. Decoding Cryptococcus: from African biodiversity to worldwide prevalence. PLoS Pathog. 2025;21(2):e1012876. pmid:39899605
  51. 51. Chen R, Jiang Y-M, Wei S-C, Wang Q-M. Kwoniella shandongensis sp. nov., a basidiomycetous yeast isolated from soil and bark from an apple orchard. Int J Syst Evol Microbiol. 2012;62(Pt 11):2774–7. pmid:22447698
  52. 52. Statzell-Tallman A, Belloch C, Fell JW. Kwoniella mangroviensis gen. nov., sp.nov. (Tremellales, Basidiomycota), a teleomorphic yeast from mangrove habitats in the Florida Everglades and Bahamas. FEMS Yeast Res. 2008;8(1):103–13. pmid:17961172
  53. 53. Wang L, Zhai B, Lin X. The link between morphotype transition and virulence in Cryptococcus neoformans. PLoS Pathog. 2012;8(6):e1002765. pmid:22737071
  54. 54. Heitman J, Carter DA, Dyer PS, Soll DR. Sexual reproduction of human fungal pathogens. Cold Spring Harb Perspect Med. 2014;4(8):a019281. pmid:25085958
  55. 55. Zhao Y, Lin X. Cryptococcus neoformans: sex, morphogenesis, and virulence. Infect Genet Evol. 2021;89:104731. pmid:33497839
  56. 56. Phadke SS, Feretzaki M, Heitman J. Unisexual reproduction enhances fungal competitiveness by promoting habitat exploration via hyphal growth and sporulation. Eukaryot Cell. 2013;12(8):1155–9. pmid:23794511
  57. 57. Fu C, Thielhelm TP, Heitman J. Unisexual reproduction promotes competition for mating partners in the global human fungal pathogen Cryptococcus deneoformans. PLoS Genet. 2019;15(9):e1008394. pmid:31536509
  58. 58. Wang L, Lin X. The morphotype heterogeneity in Cryptococcus neoformans. Curr Opin Microbiol. 2015;26:60–4. pmid:26094087
  59. 59. Fu MS, Liporagi-Lopes LC, Dos Santos SR Júnior, Tenor JL, Perfect JR, Cuomo CA, et al. Amoeba predation of Cryptococcus neoformans results in pleiotropic changes to traits associated with virulence. mBio. 2021;12(2):e00567-21. pmid:33906924
  60. 60. Sauters TJC, Roth C, Murray D, Sun S, Floyd Averette A, Onyishi CU, et al. Amoeba predation of Cryptococcus: a quantitative and population genomic evaluation of the accidental pathogen hypothesis. PLoS Pathog. 2023;19(11):e1011763. pmid:37956179
  61. 61. Idnurm A. Isolation of a fungal calcineurin A mutant suggests that amoebae can counter-select virulence attributes of microbes. Med Mycol. 2023;61(2):myad013. pmid:36708172
  62. 62. Price CTD, Hanford HE, Al-Quadan T, Santic M, Shin CJ, Da’as MSJ, et al. Amoebae as training grounds for microbial pathogens. mBio. 2024;15(8):e0082724. pmid:38975782
  63. 63. Velagapudi R, Hsueh Y-P, Geunes-Boyer S, Wright JR, Heitman J. Spores as infectious propagules of Cryptococcus neoformans. Infect Immun. 2009;77(10):4345–55. pmid:19620339
  64. 64. Walsh NM, Botts MR, McDermott AJ, Ortiz SC, Wüthrich M, Klein B, et al. Infectious particle identity determines dissemination and disease outcome for the inhaled human fungal pathogen Cryptococcus. PLoS Pathog. 2019;15(6):e1007777. pmid:31247052
  65. 65. Ortiz SC, Hull CM. Biogenesis, germination, and pathogenesis of Cryptococcus spores. Microbiol Mol Biol Rev. 2024;88(1):e0019623. pmid:38440970
  66. 66. Guerreiro MA, Springer DJ, Rodrigues JA, Rusche LN, Findley K, Heitman J, et al. Molecular and genetic evidence for a tetrapolar mating system in the basidiomycetous yeast Kwoniella mangrovensis and two novel sibling species. Eukaryot Cell. 2013;12(5):746–60. pmid:23524993
  67. 67. Metin B, Findley K, Heitman J. The mating type locus (MAT) and sexual reproduction of Cryptococcus heveanensis: insights into the evolution of sex and sex-determining chromosomal regions in fungi. PLoS Genet. 2010;6(5):e1000961. pmid:20502678
  68. 68. Cheng T, Veselská T, Křížková B, Švec K, Havlíček V, Stadler M, et al. Insight into the genomes of dominant yeast symbionts of European spruce bark beetle, Ips typographus. Front Microbiol. 2023;14:1108975. pmid:37077248
  69. 69. Basson JR, Roets F, Wingfield MJ, Aylward J. Bark beetles and their associated fungi infesting native Widdringtonia species in the Western Cape province of South Africa. Afr Entomol. 2024;32:1–8.
  70. 70. Crous PW, Luangsa-Ard JJ, Wingfield MJ, Carnegie AJ, Hernández-Restrepo M, Lombard L, et al. Fungal Planet description sheets: 785-867. Persoonia. 2018;41:238–417. pmid:30728607
  71. 71. Li A-H, Yuan F-X, Groenewald M, Bensch K, Yurkov AM, Li K, et al. Diversity and phylogeny of basidiomycetous yeasts from plant leaves and soil: proposal of two new orders, three new families, eight new genera and one hundred and seven new species. Stud Mycol. 2020;96:17–140. pmid:32206137
  72. 72. Leufvén A, Bergström G, Falsen E. Interconversion of verbenols and verbenone by identified yeasts isolated from the spruce bark beetleIps typographus. J Chem Ecol. 1984;10(9):1349–61. pmid:24317586
  73. 73. Yurkov AM, Röhl O, Pontes A, Carvalho C, Maldonado C, Sampaio JP. Local climatic conditions constrain soil yeast diversity patterns in Mediterranean forests, woodlands and scrub biome. FEMS Yeast Res. 2016;16(1):fov103. pmid:26568202
  74. 74. Kwon-Chung KJ. A new genus, filobasidiella, the perfect state of Cryptococcus neoformans. Mycologia. 1975;67(6):1197–200. pmid:765816
  75. 75. Kwon-Chung KJ. A new species of Filobasidiella, the sexual state of Cryptococcus neoformans B and C serotypes. Mycologia. 1976;68(4):943–6. pmid:790173
  76. 76. Kwon-Chung KJ. Morphogenesis of Filobasidiella neoformans, the sexual state of Cryptococcus neoformans. Mycologia. 1976;68(4):821–33. pmid:790172
  77. 77. Rodriguez-Carres M, Findley K, Sun S, Dietrich FS, Heitman J. Morphological and genomic characterization of Filobasidiella depauperata: a homothallic sibling species of the pathogenic Cryptococcus species complex. PLoS One. 2010;5(3):e9620. pmid:20224779
  78. 78. Findley K, Sun S, Fraser JA, Hsueh Y-P, Averette AF, Li W, et al. Discovery of a modified tetrapolar sexual cycle in Cryptococcus amylolentus and the evolution of MAT in the Cryptococcus species complex. PLoS Genet. 2012;8(2):e1002528. pmid:22359516
  79. 79. Sun S, Priest SJ, Heitman J. Cryptococcus neoformans mating and genetic crosses. Curr Protoc Microbiol. 2019;53(1):e75. pmid:30661293
  80. 80. Raper J, Flexer A. Mating systems and evolution of the basidiomycetes. Evolution in the higher basidiomycetes. Knoxville, TN: University of Tennessee Press. 1971. p. 149–67.
  81. 81. Fraser JA, Hsueh YP, Findley KM, Heitman J. Evolution of the mating-type locus: the basidiomycetes; 2007.
  82. 82. Phadke SS, Feretzaki M, Clancey SA, Mueller O, Heitman J. Unisexual reproduction of Cryptococcus gattii. PLoS One. 2014;9(10):e111089. pmid:25337713
  83. 83. Chang YC, Wickes BL, Miller GF, Penoyer LA, Kwon-Chung KJ. Cryptococcus neoformans STE12alpha regulates virulence but is not essential for mating. J Exp Med. 2000;191(5):871–82. pmid:10704467
  84. 84. Yue C, Cavallo LM, Alspaugh JA, Wang P, Cox GM, Perfect JR, et al. The STE12alpha homolog is required for haploid filamentation but largely dispensable for mating and virulence in Cryptococcus neoformans. Genetics. 1999;153(4):1601–15. pmid:10581270
  85. 85. Burger F, Daugeron MC, Linder P. Dbp10p, a putative RNA helicase from Saccharomyces cerevisiae, is required for ribosome biogenesis. Nucleic Acids Res. 2000;28(12):2315–23. pmid:10871363
  86. 86. Parnell KM, Bass BL. Functional redundancy of yeast proteins Reh1 and Rei1 in cytoplasmic 60S subunit maturation. Mol Cell Biol. 2009;29(14):4014–23. pmid:19433447
  87. 87. Mizushima N, Noda T, Yoshimori T, Tanaka Y, Ishii T, George MD, et al. A protein conjugation system essential for autophagy. Nature. 1998;395(6700):395–8. pmid:9759731
  88. 88. Zhu Y-H, Hyun J, Pan Y-Z, Hopper JE, Rizo J, Wu J-Q. Roles of the fission yeast UNC-13/Munc13 protein Ync13 in late stages of cytokinesis. Mol Biol Cell. 2018;29(19):2259–79. pmid:30044717
  89. 89. Wigge PA, Kilmartin JV. The Ndc80p complex from Saccharomyces cerevisiae contains conserved centromere components and has a function in chromosome segregation. J Cell Biol. 2001;152(2):349–60. pmid:11266451
  90. 90. Guterman A, Glickman MH. Complementary roles for Rpn11 and Ubp6 in deubiquitination and proteolysis by the proteasome. J Biol Chem. 2004;279(3):1729–38. pmid:14581483
  91. 91. Sridhar S, Hori T, Nakagawa R, Fukagawa T, Sanyal K. Bridgin connects the outer kinetochore to centromeric chromatin. Nat Commun. 2021;12(1):146. pmid:33420015
  92. 92. Billmyre RB, Craig CJ, Lyon J, Reichardt C, Eickbush MT, Zanders SE. Saturation transposon mutagenesis enables genome-wide identification of genes required for growth and fluconazole resistance in the human fungal pathogen Cryptococcus neoformans. bioRxiv. 2024;:2024.07.28.605507. pmid:39131341
  93. 93. Mancera E, Bourgon R, Brozzi A, Huber W, Steinmetz LM. High-resolution mapping of meiotic crossovers and non-crossovers in yeast. Nature. 2008;454(7203):479–85. pmid:18615017
  94. 94. Weber CC, Boussau B, Romiguier J, Jarvis ED, Ellegren H. Evidence for GC-biased gene conversion as a driver of between-lineage differences in avian base composition. Genome Biol. 2014;15(12):549. pmid:25496599
  95. 95. Desjardins CA, Giamberardino C, Sykes SM, Yu C-H, Tenor JL, Chen Y, et al. Population genomics and the evolution of virulence in the fungal pathogen Cryptococcus neoformans. Genome Res. 2017;27(7):1207–19. pmid:28611159
  96. 96. Lengeler KB, Wang P, Cox GM, Perfect JR, Heitman J. Identification of the MATa mating-type locus of Cryptococcus neoformans reveals a serotype A MATa strain thought to have been extinct. Proc Natl Acad Sci U S A. 2000;97(26):14455–60. pmid:11121047
  97. 97. Viviani MA, Esposto MC, Cogliati M, Montagna MT, Wickes BL. Isolation of a Cryptococcus neoformans serotype A MATa strain from the Italian environment. Med Mycol. 2001;39(5):383–6. pmid:12054047
  98. 98. Viviani MA, Nikolova R, Esposto MC, Prinz G, Cogliati M. First European case of serotype A MATa Cryptococcus neoformans infection. Emerg Infect Dis. 2003;9(9):1179–80. pmid:14528882
  99. 99. Litvintseva AP, Marra RE, Nielsen K, Heitman J, Vilgalys R, Mitchell TG. Evidence of sexual recombination among Cryptococcus neoformans serotype A isolates in sub-Saharan Africa. Eukaryot Cell. 2003;2(6):1162–8. pmid:14665451
  100. 100. Ashton PM, Thanh LT, Trieu PH, Van Anh D, Trinh NM, Beardsley J, et al. Three phylogenetic groups have driven the recent population expansion of Cryptococcus neoformans. Nat Commun. 2019;10(1):2035. pmid:31048698
  101. 101. Yadav V, Sun S, Heitman J. Uniparental nuclear inheritance following bisexual mating in fungi. Elife. 2021;10:e66234. pmid:34338631
  102. 102. Hsueh Y-P, Idnurm A, Heitman J. Recombination hotspots flank the Cryptococcus mating-type locus: implications for the evolution of a fungal sex chromosome. PLoS Genet. 2006;2(11):e184. pmid:17083277
  103. 103. Loftus BJ, Fung E, Roncaglia P, Rowley D, Amedeo P, Bruno D, et al. The genome of the basidiomycetous yeast and human pathogen Cryptococcus neoformans. Science. 2005;307(5713):1321–4. pmid:15653466
  104. 104. Huang J, Larmore CJ, Priest SJ, Xu Z, Dietrich FS, Yadav V, et al. Distinct evolutionary trajectories following loss of RNA interference in Cryptococcus neoformans. Proc Natl Acad Sci U S A. 2024;121(47):e2416656121. pmid:39536081
  105. 105. Drillon G, Carbone A, Fischer G. SynChro: a fast and easy tool to reconstruct and visualize synteny blocks along eukaryotic chromosomes. PLoS One. 2014;9(3):e92621. pmid:24651407
  106. 106. D’Souza CA, Kronstad JW, Taylor G, Warren R, Yuen M, Hu G, et al. Genome variation in Cryptococcus gattii, an emerging pathogen of immunocompetent hosts. mBio. 2011;2(1):e00342-10. pmid:21304167
  107. 107. Morrow CA, Lee IR, Chow EWL, Ormerod KL, Goldinger A, Byrnes EJ 3rd, et al. A unique chromosomal rearrangement in the Cryptococcus neoformans var. grubii type strain enhances key phenotypes associated with virulence. mBio. 2012;3(2):e00310-11. pmid:22375073
  108. 108. Yadav V, Sun S, Coelho MA, Heitman J. Centromere scission drives chromosome shuffling and reproductive isolation. Proc Natl Acad Sci U S A. 2020;117(14):7917–28. pmid:32193338
  109. 109. Gomez-Zapata PA, Tellgren-Roth C, Samils B, Stenlid J, Kaitera J, Brandstrom Durling M, et al. Mating type gene divergence is associated with life cycle differentiation in scots pine blister rust. Cold Spring Harbor Laboratory. 2025. doi: https://doi.org/10.1101/2025.05.02.651906
  110. 110. Urban M, Kahmann R, Bölker M. The biallelic a mating type locus of Ustilago maydis: remnants of an additional pheromone gene indicate evolution from a multiallelic ancestor. Mol Gen Genet. 1996;250(4):414–20. doi: https://doi.org/10.1007/BF02174029 pmid:8602158
  111. 111. Fedler M, Luh K-S, Stelter K, Nieto-Jacobo F, Basse CW. The a2 mating-type locus genes lga2 and rga2 direct uniparental mitochondrial DNA (mtDNA) inheritance and constrain mtDNA recombination during sexual development of Ustilago maydis. Genetics. 2009;181(3):847–60. pmid:19104076
  112. 112. Coelho MA, Gonçalves P, Sampaio JP. Evidence for maintenance of sex determinants but not of sexual stages in red yeasts, a group of early diverged basidiomycetes. BMC Evol Biol. 2011;11:249. pmid:21880139
  113. 113. Huang J, Hu P, Ye L, Shen Z, Chen X, Liu F, et al. Pan-drug resistance and hypervirulence in a human fungal pathogen are enabled by mutagenesis induced by mammalian body temperature. Nat Microbiol. 2024;9(7):1686–99. pmid:38898217
  114. 114. Sperschneider J, Chen J, Anderson C, Morin E, Zhang X, Lewis D, et al. A chromosome-scale genome assembly of the flax rust fungus reveals the two unusually large effector proteins, AvrM3 and AvrN. Mol Plant Microbe Interact. 2025;:10.1094/MPMI-04-25-0047-R. pmid:40435161
  115. 115. Jay P, Jeffries D, Hartmann FE, Véber A, Giraud T. Why do sex chromosomes progressively lose recombination?. Trends Genet. 2024;40(7):564–79. pmid:38677904
  116. 116. Xu L, Petit E, Hood ME. Variation in mate-recognition pheromones of the fungal genus Microbotryum. Heredity (Edinb). 2016;116(1):44–51. pmid:26306729
  117. 117. Lucotte EA, Jay P, Rougemont Q, Boyer L, Cornille A, Snirc A, et al. Repeated loss of function at HD mating-type genes and of recombination in anther-smut fungi. Nat Commun. 2025;16(1):4962. pmid:40436846
  118. 118. Shen W-C, Davidson RC, Cox GM, Heitman J. Pheromones stimulate mating and differentiation via paracrine and autocrine signaling in Cryptococcus neoformans. Eukaryot Cell. 2002;1(3):366–77. pmid:12455985
  119. 119. Zahavi A. Mate selection-a selection for a handicap. J Theor Biol. 1975;53(1):205–14. pmid:1195756
  120. 120. Nieuwenhuis BPS, Aanen DK. Sexual selection in fungi. J Evol Biol. 2012;25(12):2397–411. pmid:23163326
  121. 121. Jackson CL, Hartwell LH. Courtship in S. cerevisiae: both cell types choose mating partners by responding to the strongest pheromone signal. Cell. 1990;63(5):1039–51. pmid:2257622
  122. 122. Rogers DW, Greig D. Experimental evolution of a sexually selected display in yeast. Proc Biol Sci. 2009;276(1656):543–9. pmid:18842545
  123. 123. Jackson JA, Fink GR. Gene conversion between duplicated genetic elements in yeast. Nature. 1981;292(5821):306–11. pmid:6265790
  124. 124. Szostak JW, Orr-Weaver TL, Rothstein RJ, Stahl FW. The double-strand-break repair model for recombination. Cell. 1983;33(1):25–35. pmid:6380756
  125. 125. Hawley RS. The human Y chromosome: rumors of its death have been greatly exaggerated. Cell. 2003;113(7):825–8. pmid:12837241
  126. 126. Rozen S, Skaletsky H, Marszalek JD, Minx PJ, Cordum HS, Waterston RH, et al. Abundant gene conversion between arms of palindromes in human and ape Y chromosomes. Nature. 2003;423(6942):873–6. pmid:12815433
  127. 127. Connallon T, Clark AG. Gene duplication, gene conversion and the evolution of the Y chromosome. Genetics. 2010;186(1):277–86. pmid:20551442
  128. 128. Corcoran P, Anderson JL, Jacobson DJ, Sun Y, Ni P, Lascoux M, et al. Introgression maintains the genetic integrity of the mating-type determining chromosome of the fungus Neurospora tetrasperma. Genome Res. 2016;26(4):486–98. pmid:26893460
  129. 129. Janbon G, Ormerod KL, Paulet D, Byrnes EJ 3rd, Yadav V, Chatterjee G, et al. Analysis of the genome and transcriptome of Cryptococcus neoformans var. grubii reveals complex RNA expression and microevolution leading to virulence attenuation. PLoS Genet. 2014;10(4):e1004261. pmid:24743168
  130. 130. Yadav V, Sun S, Billmyre RB, Thimmappa BC, Shea T, Lintner R, et al. RNAi is a critical determinant of centromere evolution in closely related fungi. Proc Natl Acad Sci U S A. 2018;115(12):3108–13. pmid:29507212
  131. 131. Priest SJ, Yadav V, Roth C, Dahlmann TA, Kück U, Magwene PM, et al. Uncontrolled transposition following RNAi loss causes hypermutation and antifungal drug resistance in clinical isolates of Cryptococcus neoformans. Nat Microbiol. 2022;7(8):1239–51. pmid:35918426
  132. 132. Lemoine FJ, Degtyareva NP, Lobachev K, Petes TD. Chromosomal translocations in yeast induced by low levels of DNA polymerase a model for chromosome fragile sites. Cell. 2005;120(5):587–98. pmid:15766523
  133. 133. Byrnes EJ 3rd, Li W, Ren P, Lewit Y, Voelz K, Fraser JA, et al. A diverse population of Cryptococcus gattii molecular type VGIII in southern Californian HIV/AIDS patients. PLoS Pathog. 2011;7(9):e1002205. pmid:21909264
  134. 134. Gostinčar C, Sun X, Zajc J, Fang C, Hou Y, Luo Y, et al. Population genomics of an obligately halophilic basidiomycete Wallemia ichthyophaga. Front Microbiol. 2019;10:2019. pmid:31551960
  135. 135. Lee N, Bakkeren G, Wong K, Sherwood JE, Kronstad JW. The mating-type and pathogenicity locus of the fungus Ustilago hordei spans a 500-kb region. Proc Natl Acad Sci U S A. 1999;96(26):15026–31. pmid:10611332
  136. 136. Liang S-W, Huang Y-H, Chiu J-Y, Tseng H-W, Huang J-H, Shen W-C. The smut fungus Ustilago esculenta has a bipolar mating system with three idiomorphs larger than 500 kb. Fungal Genet Biol. 2019;126:61–74. pmid:30794950
  137. 137. Sun Y, Svedberg J, Hiltunen M, Corcoran P, Johannesson H. Large-scale suppression of recombination predates genomic rearrangements in Neurospora tetrasperma. Nat Commun. 2017;8(1):1140. pmid:29074958
  138. 138. Heitman J, Sun S, James TY. Evolution of fungal sexual reproduction. Mycologia. 2013;105(1):1–27. pmid:23099518
  139. 139. Hartmann FE, Rodríguez de la Vega RC, Carpentier F, Gladieux P, Cornille A, Hood ME, et al. Understanding adaptation, coevolution, host specialization, and mating system in castrating anther-smut fungi by combining population and comparative genomics. Annu Rev Phytopathol. 2019;57:431–57. pmid:31337277
  140. 140. Taylor JW, Hann-Soden C, Branco S, Sylvain I, Ellison CE. Clonal reproduction in fungi. Proc Natl Acad Sci U S A. 2015;112(29):8901–8. pmid:26195774
  141. 141. Koren S, Walenz BP, Berlin K, Miller JR, Bergman NH, Phillippy AM. Canu: scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation. Genome Res. 2017;27(5):722–36. pmid:28298431
  142. 142. Walker BJ, Abeel T, Shea T, Priest M, Abouelliel A, Sakthikumar S, et al. Pilon: an integrated tool for comprehensive microbial variant detection and genome assembly improvement. PLoS One. 2014;9(11):e112963. pmid:25409509
  143. 143. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009;25(14):1754–60. pmid:19451168
  144. 144. Li H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics. 2018;34(18):3094–100. pmid:29750242
  145. 145. Robinson JT, Thorvaldsdóttir H, Winckler W, Guttman M, Lander ES, Getz G, et al. Integrative genomics viewer. Nat Biotechnol. 2011;29(1):24–6. pmid:21221095
  146. 146. Gnerre S, Maccallum I, Przybylski D, Ribeiro FJ, Burton JN, Walker BJ, et al. High-quality draft assemblies of mammalian genomes from massively parallel sequence data. Proc Natl Acad Sci U S A. 2011;108(4):1513–8. pmid:21187386
  147. 147. Brůna T, Hoff KJ, Lomsadze A, Stanke M, Borodovsky M. BRAKER2: automatic eukaryotic genome annotation with GeneMark-EP+ and AUGUSTUS supported by a protein database. NAR Genom Bioinform. 2021;3(1):lqaa108. pmid:33575650
  148. 148. Goodwin TJ, Poulter RT. The diversity of retrotransposons in the yeast Cryptococcus neoformans. Yeast. 2001;18(9):865–80. pmid:11427969
  149. 149. Cantalapiedra CP, Hernández-Plaza A, Letunic I, Bork P, Huerta-Cepas J. eggNOG-mapper v2: functional annotation, orthology assignments, and domain prediction at the metagenomic scale. Mol Biol Evol. 2021;38(12):5825–9. pmid:34597405
  150. 150. Blin K, Shaw S, Kloosterman AM, Charlop-Powers Z, van Wezel GP, Medema MH, et al. antiSMASH 6.0: improving cluster detection and comparison capabilities. Nucleic Acids Res. 2021;49(W1):W29–35. pmid:33978755
  151. 151. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80. pmid:23329690
  152. 152. Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics. 2009;25(15):1972–3. pmid:19505945
  153. 153. Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol. 2020;37(5):1530–4. pmid:32011700
  154. 154. Zhang C, Rabiee M, Sayyari E, Mirarab S. ASTRAL-III: polynomial time species tree reconstruction from partially resolved gene trees. BMC Bioinformatics. 2018;19(Suppl 6):153. pmid:29745866
  155. 155. Chernomor O, von Haeseler A, Minh BQ. Terrace aware data structure for phylogenomic inference from supermatrices. Syst Biol. 2016;65(6):997–1008. pmid:27121966
  156. 156. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14(6):587–9. pmid:28481363
  157. 157. Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS. UFBoot2: improving the ultrafast bootstrap approximation. Mol Biol Evol. 2018;35(2):518–22. pmid:29077904
  158. 158. Sullivan MJ, Petty NK, Beatson SA. Easyfig: a genome comparison visualizer. Bioinformatics. 2011;27(7):1009–10. pmid:21278367
  159. 159. Ianiri G, Fang YF, Dahlmann TA, Clancey SA, Janbon G, Kück U, et al. Mating-type-specific ribosomal proteins control aspects of sexual reproduction in Cryptococcus neoformans. Genetics. 2020;214(3):635–49. pmid:31882399
  160. 160. Bian Z, Xu Z, Peer A, Choi Y, Priest SJ, Akritidou K, et al. Essential genes encoded by the mating-type locus of the human fungal pathogen Cryptococcus neoformans. bioRxiv. 2024;:2024.12.02.626420. pmid:39677606
  161. 161. Hershberg R, Petrov DA. General rules for optimal codon choice. PLoS Genet. 2009;5(7):e1000556. pmid:19593368
  162. 162. LaBella AL, Opulente DA, Steenwyk JL, Hittinger CT, Rokas A. Signatures of optimal codon usage in metabolic genes inform budding yeast ecology. PLoS Biol. 2021;19(4):e3001185. pmid:33872297
  163. 163. Wright F. The “effective number of codons” used in a gene. Gene. 1990;87(1):23–9. pmid:2110097
  164. 164. Cabanettes F, Klopp C. D-GENIES: dot plot large genomes in an interactive, efficient and simple way. PeerJ. 2018;6:e4958. pmid:29888139
  165. 165. Langdon QK, Peris D, Kyle B, Hittinger CT. sppIDer: a species identification tool to investigate hybrid genomes with high-throughput sequencing. Mol Biol Evol. 2018;35(11):2835–49. pmid:30184140
  166. 166. Fu C, Heitman J. PRM1 and KAR5 function in cell-cell fusion and karyogamy to drive distinct bisexual and unisexual cycles in the Cryptococcus pathogenic species complex. PLoS Genet. 2017;13(11):e1007113. pmid:29176784
  167. 167. Priest SJ, Coelho MA, Mixão V, Clancey SA, Xu Y, Sun S, et al. Factors enforcing the species boundary between the human pathogens Cryptococcus neoformans and Cryptococcus deneoformans. PLoS Genet. 2021;17(1):e1008871. pmid:33465111