Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jan 10.
Published in final edited form as: Nature. 2010 Dec 16;468(7326):959–963. doi: 10.1038/nature09560

Intercalation of a new tier of transcription regulation into an ancient circuit

Lauren N Booth 1, Brian B Tuch 1,, Alexander D Johnson 1
PMCID: PMC3254258  NIHMSID: NIHMS266815  PMID: 21164485

Abstract

Changes in gene regulatory networks are a major source of evolutionary novelty13. Here we describe a specific type of network rewiring event, one that intercalates a new level of transcriptional control into an ancient circuit. We deduce that, over evolutionary time, the direct ancestral connections between a regulator and its target genes were broken and replaced by indirect connections, preserving the overall logic of the ancestral circuit but producing a new behaviour. The example was uncovered through a series of experiments in three ascomycete yeasts: the bakers’ yeast Saccharomyces cerevisiae, the dairy yeast Kluyveromyces lactis and the human pathogen Candida albicans. All three species have three cell types: two mating-competent cell forms (a and α) and the product of their mating (a/α), which is mating-incompetent. In the ancestral mating circuit, two homeodomain proteins, Mata1 and Matα2, form a heterodimer that directly represses four genes that are expressed only in a and α cells and are required for mating46. In a relatively recent ancestor of K. lactis, a reorganization occurred. The Mata1–Matα2 heterodimer represses the same four genes (known as the core haploid-specific genes) but now does so indirectly through an intermediate regulatory protein, Rme1. The overall logic of the ancestral circuit is preserved (haploid-specific genes ON in a and α cells and OFF in a/α cells), but a new phenotype was produced by the rewiring: unlike S. cerevisiae and C. albicans, K. lactis integrates nutritional signals, by means of Rme1, into the decision of whether or not to mate.


In S. cerevisiae, K. lactis and C. albicans, three cell types (a, α and a/α) are specified by transcriptional regulators (sequence-specific DNA-binding proteins) encoded at the mating type locus. An important part of this cell-type-specific circuit is the regulation of the haploid-specific genes (hsgs), a group of genes that are expressed in a and α cells but not in a/α cells7. The full sets of hsgs were previously identified in S. cerevisiae4 and C. albicans (ref. 5, and B.B.T., Q. M. Mitrovich, F. M. De La Vega, C. K. Monighetti and A.D.J., unpublished observations) but not in the related species K. lactis. To examine the evolution of this portion of the mating circuit, we identified the genes in the K. lactis hsg regulon and compared them to those in S. cerevisiae and C. albicans. By profiling the expression patterns of wild-type a, α and aK. lactis cells genome-wide, we identified 12 genes that are clear hsgs under the conditions tested (Fig. 1a), two of which—RME1 (referred to previously as MTS1) and STE4—were previously identified as hsgs in K. lactis8. Comparison of all the hsgs in the three species revealed a substantial level of turnover in the regulon; in other words, an hsg in one species is not necessarily an hsg in the other two (Fig. 1b). However, an ancestral core of four hsgs (GPA1, STE4, STE18 and FAR1) share a common expression pattern in all three species. The first three genes encode the heterotrimeric G protein that, in the presence of mating pheromone, activates a downstream mitogen-activated protein kinase (MAPK) cascade required for mating911. Far1 lies further downstream in this pathway and mediates two responses needed as a prelude to mating, cell cycle arrest12 and the formation of mating projections13,14.

Figure 1. The core hsgs are not directly regulated by a1–α2 in K. lactis.

Figure 1

a, The expression profiles of the set of 12 hsgs identified in K. lactis. Note that phosphate starvation induces expression of the hsgs and is required to identify these genes. For example, when starved for phosphate, the heterotrimeric G protein genes are expressed in a and α cells at levels about fivefold higher than in a/α cells. b, A comparison of hsgs defined by transcriptional profiling in S. cerevisiae (Sc)4, K. lactis (Kl) (panel a) and C. albicans (Ca) (ref. 5, and B.B.T., Q. M. Mitrovich, F. M. De La Vega, C. K. Monighetti and A.D.J., unpublished observations) shows a conserved subset of hsgs (GPA1, STE4, STE18, FAR1), which we refer to as the core hsgs (bold in a and b). c, ChIP enrichment profiles from experiments using haemagglutinin (HA)-tagged MATa1 a/α cells (magenta), HA-tagged MATα2 a/α cells (blue) and, as a control, untagged a/α cells (green). The ChIP enrichment was determined by hybridization to a tiling microarray. The location of the a1–α2 motif in the RME1 promoter is indicated by the orange star. The genes (tan boxes) are all transcribed in the reverse direction. Data were visualized with MochiView30. d, The K. lactis a1–α2 motif determined from the ChIP-chip data. For comparison, the S. cerevisiae and C. albicans motifs (derived from published ChIP data4,6) are also shown.

In S. cerevisiae and C. albicans a/α cells, all four genes of the hsg core regulon are directly repressed by the transcription regulator a1–α2 (refs 4, 6), a heterodimer encoded by one gene at the MATa locus and one gene at the MATα locus. To determine whether this was also true in K. lactis, genome-wide chromatin immunoprecipitations (ChIP-chip) of a1 and α2 were performed in K. lactis a/α cells. In total, the upstream regions of 14 genes were observed to be occupied by both a1 and α2, including the RME1 gene (Fig. 1c), which is also a1–α2 regulated in S. cerevisiae. a1 and α2 ChIP peaks were not observed at the promoters of any of the four core hsgs in K. lactis, indicating that, unlike in S. cerevisiae and C. albicans, these genes are not directly regulated by a1–α2.

To confirm the absence of direct a1–α2 regulation at K. lactis hsgs, we identified the a1–α2 recognition motif in K. lactis from the ChIP data, using a de novo motif-finding program15. The highest-scoring motif was similar to the a1–α2 motifs previously identified in S. cerevisiae16 and C. albicans17 (Fig. 1d). Indeed, the S. cerevisiae motif is efficiently recognized by the C. albicans a1–α2 protein17, confirming that key features of this sequence have remained largely unchanged in the three species. We searched the regions 2 kilobases upstream of each K. lactis core hsg for the K. lactis a1–α2 motif but did not find significant matches, confirming the absence of direct a1–α2 regulation of these genes. (Whereas the a1–α2 site upstream of RME1 had a log10-odds score of 4.98, the best matches at the core hsgs ranged from −0.70 to 0.93.) These results indicate that although the ancestral core hsg expression pattern is conserved in K. lactis, the mechanism of the regulation has changed.

To understand how the K. lactis hsgs are cell-type regulated we searched the upstream regions of the 12 genes identified as hsgs by expression array (Fig. 1a) for cis-regulatory motifs15. The second-highest ranking motif (the top-ranking motif was a repeat sequence) was found in 11 out of 12 of the promoters (Fig. 2a) and was similar to the S. cerevisiae Rme1 motif (K. lactis consensus, GAACCNMAA; S. cerevisiae consensus, GAACCTCAA18,19). This motif is also similar to, although longer than, the K. lactis Rme1 motif derived previously20. The Rme1 motif is absent from S. cerevisiae and C. albicans hsg promoters.

Figure 2. RME1 is a direct activator of hsg expression and is required for K. lactis mating.

Figure 2

a, The K. lactis Rme1 motif found by a de novo search15 of the 12 Kl hsgs and the S. cerevisiae motif derived from two experimentally characterized binding sites18,19. b, The set of 19 genes repressed twofold or greater relative to wild type when RME1 is absent and the cells are phosphate-starved. In bold are the core hsgs. c, Rme1 is a direct regulator of the core hsgs. ChIP of Rme1 was performed in K. lactis c-Myc-tagged RME1 a cells (blue and green lines, two biological replicates) and untagged, control a cells (orange line). The immunoprecipitated DNA was hybridized to a tiling microarray. The genes (tan boxes) above the line are transcribed in the forward direction and those below are transcribed in the reverse direction. The location of the Kl Rme1 motif is indicated by a purple star. d, Rme1 is required only in K. lactis to respond to mating pheromone. Wild-type or RME1 knockout a cells were exposed to either α mating pheromone or a mock treatment of dimethylsulphoxide (DMSO). Mating projections form readily when both wild-type and Δrme1 S. cerevisiae and C. albicans cells are exposed to mating pheromone. Only the K. lactis Δrme1 a cells were unable to respond to the presence of α mating pheromone. e, Rme1 is required for mating in K. lactis but not in S. cerevisiae nor C. albicans. Quantitative mating assays were performed by mating wild-type or Δrme1 a cells to wild-type α cells. In S. cerevisiae and C. albicans the percentage of a cells that was able to mate is similar for wild-type and Δrme1. K. lactis Δrme1 a cells are mating incompetent; no mating products were isolated from the Δrme1 a × wild-type α mating.

In S. cerevisiae, Rme1 was initially identified as a repressor of meiosis and sporulation21,22, and was later shown to act as a transcriptional activator of other genes19. In K. lactis, Rme1 was shown to regulate mating-type interconversion20 (the switching of a and α cells to the opposite cell-type by means of DNA rearrangement). We speculated that Rme1 was co-opted in the K. lactis lineage to positively regulate the core hsgs.

To test this hypothesis, we knocked out the RME1 gene in K. lactis a cells and examined the gene expression profile by microarray. We found that 20 genes were downregulated in the knockout strain (Fig. 2b), including all four of the core hsgs (P = 2 × 10−10, hypergeometric distribution). We also observed a set of genes that was upregulated in the absence of Rme1 (Supplementary Fig. 1), including a significant number of genes orthologous to S. cerevisiae sporulation genes (Gene Ontology (GO): 0043934, n = 14, P = 10−14, hypergeometric distribution) indicating that the function of Rme1 in regulating meiosis is shared with S. cerevisiae. Thus, whereas some of the targets of Rme1 have remained the same as in the common ancestor of S. cerevisiae and K. lactis, Rme1 has gained new targets in the K. lactis lineage, including the core hsgs.

To test whether Rme1 directly regulates the core hsgs in K. lactis, we performed a genome-wide ChIP of Rme1 in a cells. ChIP peaks were observed at the promoters of the four core hsgs (Fig. 2c) and were centred over the Rme1 motifs (Fig. 2c). Thus, in the K. lactis lineage, Rme1 was gained as a direct activator of the core hsgs by the acquisition of Rme1 cis-regulatory sequences at all four genes. We note that Rme1 is not the only regulator of the K. lactis hsgs; for example, STE18 is repressed by Sir2 (ref. 23).

We next tested the biological role of Rme1 in mating in K. lactis, S. cerevisiae and C. albicans by comparing wild-type and RME1 knockout a cells. In response to α pheromone, a cells form mating projections (polarized growth towards the source of pheromone). When S. cerevisiae and C. albicans wild-type and Δrme1 a cells were exposed to α mating pheromone, both strains formed mating projections normally (Fig. 2d). In contrast, whereas K. lactis wild-type a cells produced mating projections in response to pheromone, Δrme1 a cells did not, indicating that this biological response was dependent on Rme1 (Fig. 2d). As a second test of the role of Rme1, we examined mating directly using a quantitative mating assay. No difference was observed between the mating efficiencies of wild-type a cells and those of Δrme1 a cells for S. cerevisiae and C. albicans (Fig. 2e). In contrast, the K. lactis Δrme1 a cell mating efficiency was decreased, relative to the wild type, by a factor of at least 106 (Fig. 2e). Thus, the ability to mate is critically dependent on Rme1—but only in K. lactis.

Unlike S. cerevisiae and C. albicans, K. lactis requires a starvation signal to mate24 and to respond to pheromone25. Although several different types of starvation signal can prime K. lactis to respond to pheromone24,25, we found that phosphate starvation is particularly potent, and it was used in subsequent experiments. Our expression profiling experiments (Fig. 1a) revealed that K. lactis requires starvation to express most of its mating genes. RME1 was also highly induced (24-fold) by phosphate starvation (Fig. 1a). We note that S. cerevisiae RME1 transcript levels also increase tenfold under starvation conditions18, suggesting that regulation of RME1 by starvation may be ancestral to S. cerevisiae and K. lactis.

We next investigated in greater detail how the starvation signal is incorporated in the K. lactis mating regulatory circuit. The simplest model consistent with the data presented so far is that starvation upregulates RME1, which in turn activates transcription of the hsgs. A prediction of this model is that ectopic expression of RME1 in K. lactis should override the requirement for starvation in expressing the hsgs. We created an a strain overexpressing RME1 to levels that were within tenfold of the level in starved wild-type cells (using the Kl LAC4 promoter) and found that overexpression of RME1 (in rich YEP-galactose medium) was sufficient to induce expression of the heterotrimeric G protein subunits (Fig. 3a). Overexpression of RME1 is also sufficient to allow K. lactis to form mating projections in response to pheromone in rich medium (Fig. 3b). These results strongly support the model by showing that upregulation of RME1 is sufficient to cause biologically relevant upregulation of the heterotrimeric G proteins.

Figure 3. Overexpression of RME1 is sufficient for hsg expression in the absence of nutrient starvation.

Figure 3

a, In the overexpression strain (pLAC4-RME1), RME1 transcription is induced by galactose-containing medium, a condition that does not cause expression of the heterotrimeric G proteins or pheromone response in wild-type (WT) cells. A strain using the empty pLAC4 vector was used as a control. The transcripts were measured relative to ACT1 transcript levels by RT–quantitative PCR (means and s.d., n = 3). In the absence of a starvation signal the hsgs, but not CKB1 (a non-hsg control), are upregulated when RME1 is overexpressed. b, RME1 overexpression allows cells to respond to mating pheromone in the absence of a starvation signal. K. lactis a cells that contained only the endogenous RME1 copy and an empty pLAC4 vector (WT), the endogenous copy of RME1 and RME1 driven by the pLAC4 promoter (WT + pLAC4-RME1) or only RME1 driven by the pLAC4 promoter (Δrme1 1pLAC4-RME1) were grown in YEP-galactose and exposed to α mating pheromone. Wild-type cells were unable to form mating projections in the absence of a starvation signal, but both strains overexpressing RME1 (pLAC4-RME1) formed mating projections in the absence of a starvation signal.

Thus, the rewiring of the K. lactis hsg circuit (summarized in Fig. 4) resulted in a new network configuration and a novel phenotype, relative to the ancestor. Our results suggest a possible evolutionary path for this rewiring. In the ancestor of all three yeasts, the hsgs were directly repressed by a1–α2. Either in an ancestor to S. cerevisiae and K. lactis or independently in each lineage, RME1 was brought under nutritional regulation. Finally, in the K. lactis lineage alone, two steps occurred: the hsgs lost the cis-regulatory sequences for a1–α2 and gained the cis-regulatory sequences for Rme1. As described in Supplementary Fig. 2, it is possible to determine more precisely when the rewiring of the core hsgs occurred. We can infer that direct a1–α2 regulation of the core hsgs was probably lost several times in the ascomycete lineage, and that the K. lactis form of regulation of the hsgs probably arose after K. lactis and the closely related species L. kluyveri branched from their common ancestor.

Figure 4. A simplified model for the evolution of regulation of core hsgs in three yeasts.

Figure 4

In all three species the core hsgs are repressed by a1–α2; thus, they are ON in a and α cells and OFF in a/α cells. In S. cerevisiae and C. albicans the repression is direct (a1–α2 binds to the promoters of these genes), but in K. lactis it is indirect, through Rme1. The circuit rewiring in the K. lactis lineage has resulted in a new mating behaviour; this species is able to mate only when starved. We show that this behaviour is due to the intercalation of Rme1, which is upregulated by starvation in K. lactis.

Although we do not know whether acquisition of the K. lactis form of regulation was adaptive, this type of regulation makes logical sense given that the primary mode of growth of K. lactis is as a haploid26. The formation of spores is a strategy employed by many yeasts to survive harsh environments. For starvation to give rise to spores, K. lactis would first have to mate (to form the sporulation-competent a/α cell type), thus rationalizing the link between starvation and mating. In contrast, S. cerevisiae in the wild is typically at least diploid27 and forms spores directly in response to starvation. Thus, the coupling of mating and starvation makes conceptual sense for K. lactis in comparison with S. cerevisiae.

We have described a case in which a new tier of regulation has been intercalated into an ancient transcription circuit consisting of a regulator (a homeodomain heterodimer) and a set of target genes. This change involved breaking the original connections between the regulator and its target genes and replacing them with a more complex type of hierarchy (Fig. 4). Intercalation may be a common way in which regulatory circuits evolve. This type of ‘intercalary evolution’ was first proposed28 to account for a common origin of eyes. In a wide variety of species, the transcription regulator Pax6 lies at the top of the eye development hierarchy, and rhodopsins occupy the bottom. According to the proposal, different types of eye arose from evolutionary intercalation of a variety of regulatory and structural genes within this simple, deeply conserved, regulatory relationship. The change we describe here is less complex and provides a concrete example of evolutionary intercalation, one that is responsible for an important feature of modern mating behaviour in K. lactis. It has been known for decades that K. lactis (unlike its relatives S. cerevisiae and C. albicans) requires starvation to mate24, and we have shown that this behaviour is due to the new configuration of the K. lactis mating circuit.

METHODS SUMMARY

Gene expression array

RNA was isolated by hot phenol extraction and reverse transcribed, and the resulting complementary DNAs were coupled to Cy5. A pooled mixture of the cDNAs was coupled to Cy3 and used as a reference. Labelled cDNAs were hybridized to Agilent arrays for visualization.

ChIP-chip

ChIP experiments were performed as described previously29, with minor modifications.

Pheromone response assays

Exponential-phase cultures were exposed to 13-mer α-mating pheromone, and the formation of mating projections was monitored by microscopy.

Quantitative mating assay

a and α cultures were grown independently to exponential phase and then mixed together with α cells in fivefold excess under mating conditions. The mating products were selected for on medium that either a and a/α cells or only a/α cells could grow on, and efficiencies were calculated as efficiency = (a/α colonies)/(a and a/α colonies).

RT–quantitative PCR

Cultures were grown to exponential phase in YEP-galactose medium, and RNA was isolated by extraction with hot phenol. RNA was reverse transcribed and the cDNAs were quantified by quantitative PCR.

Acknowledgments

We thank Q. Mitrovich, O. Homann, A. Hernday, M. Miller, C. Cain, T. Sorrells and H. Madhani for helpful discussions and technical contributions; and S. Åström for generously providing the K. lactis strains used in this study. The S. cerevisiae strains were a gift from the H. Madhani and J. Li laboratories. The work was funded by grant RO1 GM037049 from the National Institutes of Health. L.N.B. is a National Science Foundation Graduate Research Fellow.

Footnotes

Full Methods and any associated references are available in the online version of the paper at www.nature.com/nature.

Supplementary Information is linked to the online version of the paper at www.nature.com/nature.

Author Contributions L.N.B. performed all experiments. L.N.B. and B.B.T. analysed data. L.N.B., B.B.T. and A.D.J. designed the study and wrote the paper.

Author Information The gene expression array data have been deposited in the NCBI Gene Expression Omnibus (GEO) under accession number GSE24874. For the ChIP-chip data the accession number is GSE25209. Reprints and permissions information is available at www.nature.com/reprints. The authors declare no competing financial interests. Readers are welcome to comment on the online version of this article at www.nature.com/nature.

References

  • 1.Carroll S. Evo-devo and an expanding evolutionary synthesis: a genetic theory of morphological evolution. Cell. 2008;134:25–36. doi: 10.1016/j.cell.2008.06.030. [DOI] [PubMed] [Google Scholar]
  • 2.Davidson EH, Erwin DH. Gene regulatory networks and the evolution of animal body plans. Science. 2006;311:796–800. doi: 10.1126/science.1113832. [DOI] [PubMed] [Google Scholar]
  • 3.Wray GA. The evolutionary significance of cis-regulatory mutations. Nature Rev Genet. 2007;8:206–216. doi: 10.1038/nrg2063. [DOI] [PubMed] [Google Scholar]
  • 4.Galgoczy DJ, et al. Genomic dissection of the cell-type-specification circuit in Saccharomyces cerevisiae. Proc Natl Acad Sci USA. 2004;101:18069–18074. doi: 10.1073/pnas.0407611102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Tsong AE, Miller MG, Raisner RM, Johnson AD. Evolution of a combinatorial transcriptional circuit: a case study in yeasts. Cell. 2003;115:389–399. doi: 10.1016/s0092-8674(03)00885-7. [DOI] [PubMed] [Google Scholar]
  • 6.Srikantha T, et al. TOS9 regulates white-opaque switching in Candida albicans. Eukaryot Cell. 2006;5:1674–1687. doi: 10.1128/EC.00252-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Herskowitz I. A regulatory hierarchy for cell specialization in yeast. Nature. 1989;342:749–757. doi: 10.1038/342749a0. [DOI] [PubMed] [Google Scholar]
  • 8.Barsoum E, Sjöstrand JOO, Aström SU. Ume6 is required for the MATa/MATα-cellular identity and transcriptional silencing in Kluyveromyces lactis. Genetics. 2010;184:999–1011. doi: 10.1534/genetics.110.114678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Herskowitz I. MAP kinase pathways in yeast: for mating and more. Cell. 1995;80:187–197. doi: 10.1016/0092-8674(95)90402-6. [DOI] [PubMed] [Google Scholar]
  • 10.Dignard D, André D, Whiteway M. Heterotrimeric G-protein subunit function in Candida albicans: both the α and β subunits of the pheromone response G protein are required for mating. Eukaryot Cell. 2008;7:1591–1599. doi: 10.1128/EC.00077-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Coria R, et al. The pheromone response pathway of Kluyveromyces lactis. FEM Yeast Res. 2006;6:336–344. doi: 10.1111/j.1567-1364.2005.00022.x. [DOI] [PubMed] [Google Scholar]
  • 12.Chang F, Herskowitz I. Identification of a gene necessary for cell cycle arrest by a negative growth factor of yeast: FAR1 is an inhibitor of a G1 cyclin, CLN2. Cell. 1990;63:999–1011. doi: 10.1016/0092-8674(90)90503-7. [DOI] [PubMed] [Google Scholar]
  • 13.Butty AC, Pryciak PM, Huang LS, Herskowitz I, Peter M. The role of Far1p in linking the heterotrimeric G protein to polarity establishment proteins during yeast mating. Science. 1998;282:1511–1516. doi: 10.1126/science.282.5393.1511. [DOI] [PubMed] [Google Scholar]
  • 14.Côte P, Whiteway M. The role of Candida albicans FAR1 in regulation of pheromone-mediated mating, gene expression and cell cycle arrest. Mol Microbiol. 2008;68:392–404. doi: 10.1111/j.1365-2958.2008.06158.x. [DOI] [PubMed] [Google Scholar]
  • 15.Bailey TL, Elkan C. Fitting a mixture model by expectation maximization to discover motifs in biopolymers. Proc Int Conf Intell Syst Mol Biol. 1994;2:28–36. [PubMed] [Google Scholar]
  • 16.Goutte C, Johnson AD. Recognition of a DNA operator by a dimer composed of two different homeodomain proteins. EMBO J. 1994;13:1434–1442. doi: 10.1002/j.1460-2075.1994.tb06397.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Hull CM, Johnson AD. Identification of a mating type-like locus in the asexual pathogenic yeast Candida albicans. Science. 1999;285:1271–1275. doi: 10.1126/science.285.5431.1271. [DOI] [PubMed] [Google Scholar]
  • 18.Covitz PA, Mitchell AP. Repression by the yeast meiotic inhibitor RME1. Genes Dev. 1993;7:1598–1608. doi: 10.1101/gad.7.8.1598. [DOI] [PubMed] [Google Scholar]
  • 19.Toone WM, et al. Rme1, a negative regulator of meiosis, is also a positive activator of G1 cyclin gene expression. EMBO J. 1995;14:5824–5832. doi: 10.1002/j.1460-2075.1995.tb00270.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Barsoum E, Martinez P, Astrom SU. α3, a transposable element that promotes host sexual reproduction. Genes Dev. 2010;24:33–44. doi: 10.1101/gad.557310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kassir Y, Simchen G. Regulation of mating and meiosis in yeast by the mating-type region. Genetics. 1976;82:187–206. doi: 10.1093/genetics/82.2.187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Mitchell AP, Herskowitz I. Activation of meiosis and sporulation by repression of the RME1 product in yeast. Nature. 1986;319:738–742. doi: 10.1038/319738a0. [DOI] [PubMed] [Google Scholar]
  • 23.Hickman MA, Rusche LN. The Sir2-Sum1 complex represses transcription using both promoter-specific and long-range mechanisms to regulate cell identity and sexual cycleinthe yeastKluyveromyceslactis. PLoS Genet. 2009;5:e1000710. doi: 10.1371/journal.pgen.1000710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Herman A. Interspecies sex-specific growth responses in Kluyveromyces. Antonie van Leeuwenhoek. 1970;36:421–425. doi: 10.1007/BF02069042. [DOI] [PubMed] [Google Scholar]
  • 25.Tuch BB, Galgoczy DJ, Hernday AD, Li H, Johnson AD. The evolution of combinatorial gene regulation in fungi. PLoS Biol. 2008;6:e38. doi: 10.1371/journal.pbio.0060038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Kurtzman CP, Fell JW. The Yeasts: A Taxonomic Study. 4. Elsevier; 2000. [Google Scholar]
  • 27.Ezov TK, et al. Molecular-genetic biodiversity in a natural population of the yeast Saccharomyces cerevisiae from ‘Evolution Canyon’: microsatellite polymorphism, ploidy and controversial sexual status. Genetics. 2006;174:1455–1468. doi: 10.1534/genetics.106.062745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gehring WJ, Ikeo K. Pax 6: mastering eye morphogenesis and eye evolution. Trends Genet. 1999;15:371–377. doi: 10.1016/s0168-9525(99)01776-x. [DOI] [PubMed] [Google Scholar]
  • 29.Nobile CJ, et al. Biofilm matrix regulation by Candida albicans Zap1. PLoS Biol. 2009;7:e1000133. doi: 10.1371/journal.pbio.1000133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Homann OR, Johnson AD. MochiView: versatile software for genome browsing and DNA motif analysis. BMC Biol. 2010;8:49. doi: 10.1186/1741-7007-8-49. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES