Briefings in Functional Genomics, 15(2), 2016, 85–94 doi: 10.1093/bfgp/elv035 Advance Access Publication Date: 30 September 2015 Review paper

Budding off: bringing functional genomics to Candida albicans Matthew Z. Anderson and Richard J. Bennett Corresponding author: Richard J. Bennett, Department of Microbiology and Immunology, Brown University, Providence, RI 02912, USA. Tel.: (401) 863-6341; Fax: (401) 863-2925; E-mail: [email protected]

Abstract Candida species are the most prevalent human fungal pathogens, with Candida albicans being the most clinically relevant species. Candida albicans resides as a commensal of the human gastrointestinal tract but is a frequent cause of opportunistic mucosal and systemic infections. Investigation of C. albicans virulence has traditionally relied on candidate gene approaches, but recent advances in functional genomics have now facilitated global, unbiased studies of gene function. Such studies include comparative genomics (both between and within Candida species), analysis of total RNA expression, and regulation and delineation of protein–DNA interactions. Additionally, large collections of mutant strains have begun to aid systematic screening of clinically relevant phenotypes. Here, we will highlight the development of functional genomics in C. albicans and discuss the use of these approaches to addressing both commensalism and pathogenesis in this species. Key words: genome plasticity; comparative genomics; transcriptional rewiring; aneuploidy

The human fungal pathogen Candida albicans is a commensal of the gut, mouth and genital tract, but is also capable of causing disease in mucosal niches or spreading to the bloodstream to cause more serious systemic infections [1]. Individuals with compromised immune systems are traditionally those primarily at risk for C. albicans pathogenesis [2–4]. In addition, the increased usage of immunosuppressive drugs, steroids and antibiotics has further amplified the prevalence of C. albicans in the clinic and its importance as a human pathogen [2, 5]. The Candida community has made great strides in defining the biological traits that contribute to both commensalism and virulence. Initial studies established that multiple traits contribute to C. albicans pathogenesis, particularly the ability to switch between yeast and hyphal forms, as well as to sense and adapt to multiple environmental cues [6–8]. To define the genetic determinants that contribute to pathogenicity, functional genomics approaches have been increasingly applied [9–11]. In particular, sequencing and annotation of the standard laboratory strain of C. albicans, SC5314, has greatly advanced the utility of ‘omics’ approaches [9, 12, 13]. In the post-genomic era, numerous techniques ranging from comparative genomics and

transcriptional profiling to construction of a complete gene deletion collection are being used to dissect key traits in C. albicans biology, as will be outlined in this review.

Comparative phylogenetic analysis of C. albicans Molecular typing of Candida species initially relied on DNA fingerprinting to differentiate strain types [14–16]. This method used variations in DNA repeat lengths to define the Candida species and strain subtype [15, 17]. However, variation in fingerprinting methods and loci led to the development of a consensus multi-locus sequence typing (MLST) scheme, in which researchers analyzed the sequences of seven genetically unlinked loci encoding 107 distinct single nucleotide polymorphisms (SNPs) [18, 19]. Subsequent MLST analysis of large numbers of isolates revealed a structured population with most strains falling into 1 of 17 clades [20, 21]. These methods have provided an overview of the phylogenetic relationship among isolates and showed C. albicans to be a largely clonal

Richard Bennett is a Professor in the Department of Molecular Microbiology and Immunology at Brown University. His interests include mechanisms of pathogenesis in human and animal fungal pathogens. Matthew Anderson is a Postdoctoral research associate in the laboratory of Dr. Richard Bennett at Brown University. His interests lie in determining the contributions of natural genomic variation to phenotypic diversity. C The Author 2015. Published by Oxford University Press. All rights reserved. For permissions, please email: [email protected] V

85

86

| Anderson and Bennett

population [20, 21]. However, MLST karyotyping relied on genetic information at a limited number of loci, which may have obfuscated identification of hybrid strains or introgressed genomic regions resulting from sexual exchange [22]. Sequencing of the C. albicans SC5314 genome established that there are eight diploid chromosomes [12] with extensive heterozygosity between chromosome homologs [23]. A finished draft of the genome [10] identified 6100 open reading frames (ORFs), including several multi-gene families related to pathogenesis. Comparison of gene content among eight Candida species revealed gene family expansions of cell surface transporters, lipases, proteases and genes associated with hyphal formation in C. albicans [9]. Subsequent comparison of C. albicans with its closest known relative, Candida dubliniensis, uncovered expansion of two gene families specifically in C. albicans, that of subtelomeric transcriptional regulators (TLOs) and transmembrane transporters (IFAs) [24]. Taken together, these studies alluded to both previously identified [25, 26] and additional expanded gene families in C. albicans, with potential implications for pathogenesis.

Comparative genomics among C. albicans isolates Comparison of a second sequenced C. albicans isolate, WO-1, with the reference strain, SC5314, revealed the strains to be largely syntenic with relatively few strain-specific genes [9]. To provide a more detailed analysis of genotypic and phenotypic diversity within C. albicans, a recent study [11] sequenced 21 clinical C. albicans isolates representing seven divergent MLST types [19]. The average nucleotide diversity between any two sequenced strains was 0.37%, with 6069 (98.1%) of the genes shared by any two isolates. Among all 21 strains, 461 genes (7.4%) were disrupted in at least 1 strain compared with 57 genes (0.86%) disrupted by nonsense mutations among 71 sequenced Saccharomyces cerevisiae and Saccharomyces paradoxus isolates [27]. Genes with the highest indicators of positive selection encoded cell wall proteins (PGAs), secreted virulence factors (SAP5 and PEP3) and regulators of filamentous growth (FGRs) [11]. The greatest variation in gene copy number was associated with repetitive elements such as the TAR1 gene within the rDNA array and the TLOs in subtelomeric regions. The sequences of the TLO homologs were also highly variable, with novel TLO clades evident that do not resemble those of previously sequenced family members [28]. TLO genes encode a single component of the Mediator transcriptional complex [29], and diversification of these paralogs may indicate specialized roles for different TLOs in modulating Mediator activity. Intraspecies comparisons between C. albicans isolates have also begun to reveal natural polymorphisms that contribute to pathogenesis. One clinical isolate, P94015, encoded a homozygous nonsense mutation in the master transcriptional regulator EFG1 that controls white-opaque switching [30, 31] and filamentation [32]. Loss of EFG1 function in P94015 decreased virulence in a systemic model of infection but enhanced fitness in a model of commensal gut colonization, similar to the phenotypes of efg1D/D mutants in the SC5314 background [33, 34]. Comparative genomics is, therefore, beginning to reveal natural genetic differences that contribute to fitness advantages in specific host niches. Application of whole-genome sequencing to Candida strains undergoing experimental evolution is still in its infancy but is already providing fascinating insights into C. albicans biology.

Engulfment of C. albicans by macrophages normally elicits a filamentation response that ultimately leads to rupture of the phagocyte and release of the ingested cells [35–37]. Transcription factor mutants defective in filamentation were co-cultured with macrophages, leading to an evolved C. albicans lineage that regained the ability to undergo filamentation. Sequencing of the evolved strain identified a single nucleotide mutation in the SSN3 gene, encoding a component of the Mediator complex, which rescued the filamentation response to macrophage internalization and circumvented the missing transcription factors [38]. This study demonstrates how central transcriptional pathways can be readily bypassed by rewiring of existing signaling networks. Genome sequencing of sequential C. albicans isolates recovered from patients undergoing antifungal treatment has established that multiple isolates can be present in a single infection, and that resident strains evolve extensively during the course of antifungal treatment [39]. In particular, loss of heterozygosity (LOH) was often observed in these isolates and was associated with increased drug resistance. Aneuploid forms also frequently arose in the population and, although not directly associated with increased drug resistance, may facilitate strain adaptation by promoting additional genomic changes.

Genomic architecture Genome plasticity is a hallmark feature of C. albicans and other fungal species that can promote adaptation to host or environmental stresses. Although C. albicans is typically isolated as a diploid, it is also capable of propagating in the haploid or tetraploid states [22, 40]. Furthermore, as noted above, aneuploid cells carrying an abnormal complement of chromosomes are frequently found among clinical C. albicans strains [11, 41–43], suggesting they can provide a selective advantage in some host contexts. Indeed, C. albicans exposure to the most frequently administered antifungal fluconazole induces aneuploidy in vitro [44], and resistance in the clinic is associated with an extra isochromosome of chromosome 5 in a subset of strains [42, 43, 45, 46]. This isochromosome encodes the molecular target of fluconazole, ERG11 [47], as well as TAC1, the transcriptional regulator of drug efflux pumps [48], and the increased copy numbers of both of these genes contribute to the increased drug resistance [43, 46]. Comparative genome hybridization (CGH) served as the first genome-wide method to identify whole chromosome and segmental aneuploidy (Figure 1) [49, 50]. The microarrays used for CGH were designed to distinguish 39 000 alleles between SC5314 chromosome homologs based on SNP differences. CGH arrays were used to determine both the chromosome copy number and the allelic composition, the latter allowing the identification of LOH events [49, 51]. LOH occurs frequently in C. albicans [11, 51–53], although less so than in diploid S. cerevisiae strains [54], and often influences clinically relevant phenotypes. For example, loss of a low-activity TAC1 allele and gain of a hyperactive allele increased fluconazole resistance in C. albicans [45, 55]. Studies of C. albicans karyotypes have transitioned to using next-generation sequencing (NGS) to assess chromosome copy number and allelic configuration. NGS offers many advantages over the SNP–CGH approach including: increased genome coverage, greater sensitivity and the ability to assess strains independent of genetic background [11, 39, 56, 57]. Techniques such as restriction-site-associated DNA sequencing (RAD-Seq)

Bringing functional genomics to Candida albicans |

87

Figure 1. Methods to detect the genomic architecture of C. albicans strains. Candida albicans karyotyping can be performed via a number of different methods including contour-clamped homogenous electric field (CHEF) electrophoresis, SNP–CGH, RAD-Seq and whole-genome sequencing (WGS). All methods are able to detect aneuploidy, although CHEF electrophoresis is often limited to atypical-sized chromosomes. Resolution has increased with advancements in genome sequencing, which provides the ability to detect allelic ratios from different strain backgrounds. (A colour version of this figure is available online at: http://bfg.oxfordjournals.org)

are also emerging, primarily because they dramatically reduce the cost of karyotyping fungal strains by NGS [58, 59]. This approach was recently applied to identifying relative chromosome copy numbers in C. albicans [60] and, similar to full genome sequencing, requires complementary analysis by DNA staining and flow cytometry to infer overall ploidy [56, 60]. Sequencing methods allow SNP frequencies to define allelic ratios and LOH events [11, 39, 40], and can be performed independently of a reference genome sequence [61]. YMAP, a recently developed informatics pipeline, facilitates this analysis for several C. albicans strain backgrounds by processing sequence reads directly and constructing a visualization of copy number variation and allelic ratios from SNP–CGH or NGS data [62].

Investigation of chromatin structure and transcriptional regulation Epigenetic switches are a critical aspect of C. albicans biology, and yet the chromatin changes that accompany such transitions have not been extensively evaluated. C. albicans modifies histones through acetylation [63–65], methylation [66], phosphorylation [67] and, interestingly, biotinylation [67], but the contribution of specific chromatin marks to gene regulation remains unclear. Nearly all studies to date have investigated chromatin modifications at select loci instead of whole-genome profiling. However, the effects of individual chromatin marks cannot be easily interpreted in isolation, and the entire repertoire of modifications should be considered at any given locus [68]. One of the few examples of genome-wide chromatin mapping is that performed on the C. albicans Set3C histone deacetylase complex [69]. This complex controls the yeastto-hyphal transition by regulating the expression levels of four transcription factors, BRG1, EFG1, NRG1 and TEC1. Chromatin immunoprecipitation and DNA sequencing (ChIP-Seq) showed that the Set3C complex is selectively recruited to the coding sequences of these regulators and downstream target genes, where it modulates their transcription kinetics and thereby controls the yeast-to-hyphal transition [69]. Epigenetic regulation also contributes to another major cell state decision, that of the white-opaque switch. C. albicans cells

can alternate between white and opaque forms, and these two heritable states show marked differences in mating, immune cell interactions and virulence [70, 71]. The master transcription factor regulating the white-opaque transition is Wor1, which ensures stable switching to the opaque state by positive feedback on its own promoter [72–74]. Switching is also regulated by acetylation of the WOR1 promoter; acetylation of histone H3 on lysine 56 (H3K56) by the fungal-specific acetyltransferase Rtt109 promotes stable formation of the opaque state [63]. Interestingly, Rtt109 also functions in protecting cells from genotoxic stresses, including free reactive oxygen species produced by macrophages [64]. However, the molecular targets of Rtt109 that contribute to stress resistance are unknown, and highlight the need for genome-wide studies to assess the function of chromatin marks as part of the ‘histone code’. In contrast to chromatin modifications, the genome-wide binding profiles of multiple C. albicans transcription factors have now been mapped either by chromatin immunoprecipitation and microarray analysis (ChIP-chip) or, more rarely, by ChIPseq. In the case of the white-opaque switch, a network of at least six transcription factors (AHR1, CZF1, EFG1, WOR1, WOR2 and WOR3) controls expression of 748 target genes [30, 75–78]. A detailed analysis of this circuit included ChIP-chip, global transcriptome analysis and mechanically induced trapping of molecular interactions (MITOMI) [75]. The latter technique involves testing the binding of an in vitro translated protein to all possible 8-mer DNA sequences to determine DNA-binding specificity [79, 80]. Using ChIP-chip, the six transcription factors regulating the white-opaque switch were found to bind to a restricted number of genomic regions (72–370), with significant overlap among the binding sites for all six proteins. These transcription factors, therefore, co-regulate target genes, as well as one other, to form a highly connected network of feedback and feedforward loops to establish the two heritable cell states. A similar interlocking network of six transcription factors was shown to regulate biofilm formation in C. albicans [81]. Biofilms are three-dimensional communities of cells surrounded by an extracellular matrix, and play important roles in device-associated infections and drug resistance by C. albicans strains [81, 82]. EFG1 is a central regulator of both the white-opaque circuit and the biofilm circuit, indicating a mechanistic connection between the two programs [83, 84]. Investigation of regulatory circuits such as these is providing

88

| Anderson and Bennett

new insights into the evolution of genetic networks, and reveals that complex networks are not necessarily a consequence of optimization by natural selection, but can instead arise via mostly nonadaptive mutations [85]. Studies in C. albicans have also investigated the mechanisms by which transcription factors can evolve to recognize new sets of target genes. For example, one recent study used ChIP-chip and MITOMI to determine the in vivo binding sites and in vitro DNA-binding specificities of four transcription factors that arose by successive duplications of an ancestral gene [86]. Functional specialization was due to a combination of mechanisms including small changes in intrinsic binding specificities, differences in half-site preferences (e.g. direct versus inverted repeats) and association with different cofactors [86]. These and other studies [87–90] are highlighting how comparative studies between C. albicans and related yeast species are addressing fundamental aspects of transcription circuit evolution.

Transcriptional profiling of C. albicans Custom microarrays were initially used to assay expression of select gene subsets following preliminary characterization of the C. albicans genome [91–93]. Completion of the genome sequence subsequently enabled global analysis of gene expression [94, 95], including the use of high-resolution tiling arrays to assay genome-wide transcription and identify noncoding RNAs and antisense transcripts [96]. Most recent studies of gene expression have used RNA sequencing (RNA-Seq) to obtain greater sensitivity and to reduce bias in assayed gene targets. RNA-Seq studies have identified 600 novel coding and over 1000 noncoding transcripts in C. albicans, many of which are regulated by environment and cell state [97, 98]. Transcriptional profiles have now been defined under a wide variety of conditions both in vitro and in vivo [99–103]. Profiling of unique phenotypic states, including white, opaque, gray and GUT (gastrointestinally induced transition), has been used to define the transcriptional networks regulating these states, and has revealed that these networks resemble those controlling cell fate in metazoans [34, 75, 98, 104]. RNA-Seq has also provided new mechanistic insights into gene regulation in C. albicans. For example, experiments have revealed the widespread presence of antisense transcripts and untranslated regions of varying lengths, dependent both on cell state and environmental conditions [97, 98, 105]. In the case of the white-opaque switch, RNA-Seq analysis of sense–antisense transcript pairs showed a significant anti-correlation, suggesting that many antisense transcripts act to repress their sense counterparts [98]. Antisense transcripts also regulate the yeastto-hyphal switch; the Brg1 transcription factor induces an antisense transcript against the NRG1 sense transcript, thereby targeting the sense transcript for destruction and stimulating hypha-specific gene expression [106]. Candida albicans is a diploid organism, with most strains showing extensive heterozygosity between chromosome homologs. This has enabled analysis of allele-specific expression (ASE) by both RNA-Seq and ribosome profiling [107, 108]. Ribosome profiling provides a quantitative measure of protein translation by determining which messenger RNA transcripts are physically protected from enzymatic digestion by ribosome binding. Experiments revealed differential expression of C. albicans alleles because of biases in both transcriptional and translational regulation of the two alleles. Interestingly, these processes often favored the same allele, thereby exacerbating differences in ASE [107].

To examine the transcriptional response of C. albicans to the host, much effort has focused on the development of in vitro models of the host environment. Interaction with both macrophages and neutrophils has highlighted C. albicans gene expression changes in metabolism, filamentation and protective enzymes for reactive oxygen species and other antimicrobial factors [109–112]. Expression profiling has also been performed using C. albicans cells grown on either reconstituted human oral epithelia or epithelial monolayers [113, 114]. Here, gene transcription reflected the various chronological stages of attachment, hyphal production, cell invasion and tissue destruction [99]. A similarly ordered transcriptional profile was also identified using a murine model of vulvovaginal candidiasis [115, 116]. More recently, RNA-Seq was used to evaluate both C. albicans and host transcriptomes during the interaction between fungal cells and human endothelial or oral epithelial cells [117]. These detailed studies implicated platelet-derived growth factor BB (PDGF BB) and neural precursor-cell-expressed developmentally down-regulated protein 9 (NEDD9) as host factors that promote the uptake of C. albicans cells by host cells. Both host pathways were also implicated in systemic infection, as PDGF-and NEDD9-pathway genes were up-regulated in murine kidneys during disseminated candidiasis [117]. Expression profiling of C. albicans cells during murine systemic infection has been facilitated by microarrays that limit cross-hybridization of host transcripts [100, 118]. Candida albicans genes differentially expressed during renal infection were associated with multiple processes including filamentation, stress response, adhesion and fatty acid metabolism. A subsequent analysis used nanoString to assay expression of 248 C. albicans genes during systemic infection [119]. The nanoString technology relies on the hybridization of RNA transcripts with specific capture and reporter probes; the reporter probe carries a fluorescent bar code for detection, while the capture probe allows the tripartite RNA/probe complex to be immobilized for imaging [120]. This provides for a highly sensitive and specific measure of RNA across a broad range of expression levels. Application of this technique to C. albicans revealed that systemic infection could be divided into early and late responses. The early response correlated with hyphal production and zinc and iron regulation, whereas the late response resembled that of cells grown in the presence of macrophages, including a response to oxidative stress [110]. A similar temporal analysis using microarrays divided C. albicans infections of zebrafish into sequential adhesion, invasion and damage phases [121]. Colonization of the host gastrointestinal (GI) tract by C. albicans has also been analyzed by expression profiling. Cells isolated from the cecum during mouse colonization expressed genes associated with both exponential and post-exponential growth [122]. Interestingly, a subsequent study [123] demonstrated that gene expression of cells in different GI niches (the cecum versus the ileum) varied as much from each other as they did from in vitro conditions. Furthermore, cells from both sites expressed hyphal-associated genes, despite growing exclusively in the yeast form [122, 123]. Taken together, these data suggest the GI tract includes diverse microenvironments that C. albicans is able to recognize and adapt to. A constant improvement in bioinformatic tools for C. albicans has provided methods for examining these large data sets. The Candida Genome Database (CGD; candidagenome.org) has played a key role in providing diverse tools for the study of C. albicans and related Candida pathogens. For example, CGD implemented Gene Ontology (GO) assignments to genes as well as tools to identify overrepresented GO terms on user-defined

Bringing functional genomics to Candida albicans |

sets of gene IDs [124]. Furthermore, expression patterns associated with genes involved in known processes have provided functional assignment of novel regulators to critical processes such as hypoxia [125], drug resistance [126] and filamentation [127]. Network analysis has allowed integration of both temporal expression data and transcription factor binding data to construct complex models of transcriptional regulation during the hypoxic response [125]. As the number of data sets increases and the complexity of the systems multiply, new programs must weigh the ability of these models to accurately depict transcriptional effects [128–130]. Progressively, integration of previously published expression data sets is informing our understanding of C. albicans transcriptional networks to create more robust and dynamic models of regulation [75, 125].

Construction of systematic strain collections Genetic analysis of C. albicans is complicated by an altered genetic code (translation of the CTG codon as serine instead of leucine) [131], the lack of a conventional sexual cycle [132] and the requirement for two rounds of disruption to make homozygous deletion mutants. Despite these constraints, concerted efforts from multiple labs have generated large strain collections of mutants for high-throughput screening (Table 1). One of the first C. albicans deletion libraries used a transposonmediated strategy involving a Tn7-UAU1 vector system. This method was used to construct 217 homozygous deletion strains [133], and identified a key regulator of alkaline-dependent filamentation [134]. This technique was subsequently used in targeted gene disruptions [135, 136] and additional deletion libraries [137–140]. A separate approach relied on the generation of heterozygous mutants by insertional mutagenesis with a URA3-marked transposon into the genome [141]. Subsequently, a heterozygous mutant library of 3633 unique strains was constructed that allowed easier identification of the disrupted locus and facilitated pooling experiments by bar coding each heterozygous mutant [142]. A further development was a library in which one allele was deleted and the second was placed under the control of a tetracycline-repressible promoter [143]. This library, known as the gene replacement and controlled expression (GRACE) collection, includes 2356 genes and has proven useful in assessing conditional lethality [143], functional conservation between S. cerevisiae and C. albicans genes [142], chemical genetic approaches for novel therapeutics [144] and gene function [145]. However, leaky gene expression at some

89

loci can complicate analysis [143] and argues for the use of complementary approaches to define function. A systematic approach to constructing a large, signaturetagged deletion collection was undertaken and comprised strains for 674 individually deleted genes [146]. Phenotyping of these strains identified 115 gene deletions with significant defects in virulence in a murine systemic model, 133 gene deletions with defects in filamentation in vitro and 68 mutants with defects in doubling times at 37 C [146]. Significant correlations existed between virulence and filamentation, but not between virulence and growth rates. Related work analyzed the function of 143 C. albicans transcription factors using a gene deletion collection [147], as well as the function of many of these transcription factors in GI colonization [148]. Currently, there are ongoing efforts to construct a C. albicans ORFeome containing a gene deletion library for all 6200 ORFs in the SC5314 genome [149], which will provide a powerful resource for genetic studies upon its release. Genetic redundancy and epistasis can often mask the effects of gene deletion on a given phenotype. Overexpression of candidate genes can therefore be used to complement studies of gene disruption [150]. A number of regulatable promoters have been developed to regulate gene expression, including the MET3 promoter [151], the MAL2 promoter [152] and tetracyclineregulated promoters (Tet-OFF and Tet-ON) [153, 154]. A library of overexpression strains was constructed in a Tet-ON system for 107 putative transcription factors, most with a known biological role [155]. Exploration of novel functions among these defined regulators identified a role for TEC1 and GAT2 in biofilm formation [155, 156]. An alternative overexpression strategy fused the Gal4 activation domain [157] to the C-terminus of 82 zinc cluster transcription factors to construct hyperactive forms of the fusion proteins [158]. The C. albicans ORFeome project is concurrently building an overexpression library for all 6200 genes under the control of the Tet-ON or PCK1 promoters [159, 160]. Future efforts in genome engineering will be greatly facilitated by the recent development of CRISPR/Cas9 for C. albicans, which allows efficient and rapid editing of individual genes and even gene families in the diploid organism [161].

Conclusions Functional genomics has provided profound insights into C. albicans pathogenesis. Fortuitously, increased access to these technologies, reduced costs and more comprehensive methods to analyze the resulting data are expected to translate to

Table 1. Mutant libraries constructed in C. albicans Library

Method

Number of mutants

Number of loci affected

Number of alleles affected

Essential genes

Davis et al. (2002) [134], Noble and Mitchell (2005) [138] and Blankenship et al. (2010) [139] Uhl et al. (2003) [141] Roemer et al. (2003) [143] Homann et al. (2009) [147] Sahni et al. (2010) [155] Oh et al. (2010) [142] Noble et al. (2010) [146] Chauvel et al. (2012) [159]

UAU method (transposon)

2000

380

Both

No

Insertional (transposon) Het. deletion and ptet replacement Targeted disruption ptet overexpression Insertional (transposon) Targeted disruption ptet overexpression/ pPCK1 overexpression ptet overexpression

18 000 2304 365 107 21 468 3000 768

Unknown 1152 166 107 3633 674 384

One Both Both One One Both One

Yes Yes No Yes Yes No Yes

82

82

One

Possibly

Schillig and Morschauser (2013) [158]

90

| Anderson and Bennett

increased usage of these approaches. Many of the most enlightening discoveries in C. albicans have combined multiple approaches to address a biological question. Combinatorial approaches will continue to require advancements in analysis and tool development and standardization of technique methodologies, as has occurred with other systems [162–164]. These measures will aid the increasing rate of discovery in Candida biology and ensure consistency between researchers in this rapidly emerging field.

Key Points

4.

5.

6.

7.

• Comparative genomics provides evidence of selective

expansion of gene families in C. albicans, as well as rapid evolution of gene products involved in host interactions. • Use of whole-genome approaches such as ChIP-chip and RNA-Seq has proven instrumental in revealing the regulatory architecture of key biological processes such as cell state switching. • Development of multiple strain deletion libraries has provided powerful genetic tools for dissection of gene functions.

8.

9.

10.

11.

Acknowledgements The authors would like to thank Iuliana Ene and Emily Mallick for comments on the article. We would like to pay special attention to the contributions of the Stanford Genome Technology Center in the original effort to sequence C. albicans, the Broad Institute in sequencing multiple Candida species and addressing genomic diversity within C. albicans and the Candida Genome Database (CGD) for the construction and maintenance of an invaluable resource to the Candida community for investigation and annotation of the C. albicans genome.

Funding This work was supported by NIH grants AI081704 and AI12363 (to R.J.B.), by the Burroughs Wellcome Fund (Investigator in Pathogenesis of Infectious Disease award to R.J.B.) and a NIH Research Supplement to Promote Diversity in Health-Related Research AI081704-S1 (to M.Z.A.).

12.

13.

14.

15.

16.

17.

References 1.

2.

3.

McCullough MJ, Ross BC, Reade PC. Candida albicans: a review of its history, taxonomy, epidemiology, virulence attributes, and methods of strain differentiation. Int J Oral Maxillofac Surg 1996;25(2):136–44. Colombo AL, Perfect J, DiNubile M, et al. Global distribution and outcomes for Candida species causing invasive candidiasis: results from an international randomized double-blind study of caspofungin versus amphotericin B for the treatment of invasive candidiasis. Eur J Clin Microbiol Infect Dis 2003;22(8):470–4. Cassone A, Cauda R. Candida and candidiasis in HIV-infected patients: where commensalism, opportunistic behavior and frank pathogenicity lose their borders. AIDS 2012;26(12):1457–72.

18.

19.

20.

Imam N, Carpenter CC, Mayer KH, et al. Hierarchical pattern of mucosal Candida infections in HIV-seropositive women. Am J Med 1990;89(2):142–6. Fukushima C, Shimoda T, Kawano T, et al. Effects of amphotericin B gargles on oral colonization of Candida albicans in asthmatic patients on steroid inhalation therapy. Respiration 2001;68(5):465–70. Balish E, Phillips AW. Growth and virulence of Candida albicans after oral inoculation in the chick with a monoflora of either Escherichia coli or Streptococcus faecalis. J Bacteriol 1966;91(5):1744–9. Rustchenko EP, Howard DH, Sherman F. Variation in assimilating functions occurs in spontaneous Candida albicans mutants having chromosomal alterations. Microbiology 1997;143(Pt 5):1765–78. Weissman Z, Berdicevsky I, Cavari BZ, et al. The high copper tolerance of Candida albicans is mediated by a P-type ATPase. Proc Natl Acad Sci USA 2000;97(7):3520–5. Butler G, Rasmussen MD, Lin MF, et al. Evolution of pathogenicity and sexual reproduction in eight Candida genomes. Nature 2009;459(7247):657–62. van het Hoog M, Rast TJ, Martchenko M, et al. Assembly of the Candida albicans genome into sixteen supercontigs aligned on the eight chromosomes. Genome Biol 2007;8(4):R52. Hirakawa MP, Martinez DA, Sakthikumar S, et al. Genetic and phenotypic intra-species variation in Candida albicans. Genome Res 2015;25(3):413–25. Jones T, Federspiel NA, Chibana H, et al. The diploid genome sequence of Candida albicans. Proc Natl Acad Sci USA 2004;101(19):7329–34. Braun BR, van Het Hoog M, d’Enfert C, et al. A human-curated annotation of the Candida albicans genome. PLoS Genet 2005;1(1):36–57. Thanos M, Schonian G, Meyer W, et al. Rapid identification of Candida species by DNA fingerprinting with PCR. J Clin Microbiol 1996;34(3):615–21. Steffan P, Vazquez JA, Boikov D, et al., Identification of Candida species by randomly amplified polymorphic DNA fingerprinting of colony lysates. J Clin Microbiol 1997;35(8):2031–9. Lockhart SR, Reed BD, Pierson CL, et al. Most frequent scenario for recurrent Candida vaginitis is strain maintenance with “substrain shuffling": demonstration by sequential DNA fingerprinting with probes Ca3, C1, and CARE2. J Clin Microbiol 1996;34(4):767–77. Pujol C, Pfaller M, Soll DR. Ca3 fingerprinting of Candida albicans bloodstream isolates from the United States, Canada, South America, and Europe reveals a European clade. J Clin Microbiol 2002;40(8):2729–40. Bougnoux ME, Tavanti A, Bouchier C, et al. Collaborative consensus for optimized multilocus sequence typing of Candida albicans. J Clin Microbiol 2003;41(11):5265–6. Tavanti A, Gow NA, Senesi S, et al. Optimization and validation of multilocus sequence typing for Candida albicans. J Clin Microbiol 2003;41(8):3765–76. Tavanti A, Davidson AD, Fordyce MJ, et al. Population structure and properties of Candida albicans, as determined by multilocus sequence typing. J Clin Microbiol 2005;43(11):5601–13.

21. Odds FC, Bougnoux ME, Shaw DJ, et al., Molecular phylogenetics of Candida albicans. Eukaryot Cell 2007;6(6):1041–52. 22. Forche A, Alby K, Schaefer D, et al. The parasexual cycle in Candida albicans provides an alternative pathway to meiosis

Bringing functional genomics to Candida albicans |

23.

24.

25.

26.

27. 28.

29.

30.

31.

32.

33.

34.

35.

36.

37. 38.

39.

40.

41.

for the formation of recombinant strains. PLoS Biol 2008;6(5):e110. Magee BB, Magee PT. Recent advances in the genomic analysis of Candida albicans. Rev Iberoam Micol 2005;22(4):187–93. Jackson AP, Gamble JA, Yeomans T, et al. Comparative genomics of the fungal pathogens Candida dubliniensis and Candida albicans. Genome Res 2009;19(12):2231–44. Hoyer LL, Scherer S, Shatzman AR, et al. Candida albicans ALS1: domains related to a Saccharomyces cerevisiae sexual agglutinin separated by a repeating motif. Mol Microbiol 1995;15(1):39–54. Magee BB, Hube B, Wright RJ, et al. The genes encoding the secreted aspartyl proteinases of Candida albicans constitute a family with at least three members. Infect Immun 1993;61(8):3240–3. Liti G, Carter DM, Moses AM, et al. Population genomics of domestic and wild yeasts. Nature 2009;458(7236):337–41. Anderson MZ, Baller JA, Dulmage K, et al. The three clades of the telomere-associated TLO gene family of Candida albicans have different splicing, localization, and expression features. Eukaryot Cell 2012;11(10):1268–75. Zhang A, Petrov KO, Hyun ER, et al. The Tlo proteins are stoichiometric components of Candida albicans mediator anchored via the Med3 subunit. Eukaryot Cell 2012;11(7):874–84. Zordan RE, Miller MG, Galgoczy DJ, et al. Interlocking transcriptional feedback loops control white-opaque switching in Candida albicans. PLoS Biol 2007;5(10):e256. Sonneborn A, Tebarth B, Ernst JF. Control of white-opaque phenotypic switching in Candida albicans by the Efg1p morphogenetic regulator. Infect Immun 1999;67(9):4655–60. Stoldt VR, Sonneborn A, Leuker CE, et al. Efg1p, an essential regulator of morphogenesis of the human pathogen Candida albicans, is a member of a conserved class of bHLH proteins regulating morphogenetic processes in fungi. EMBO J 1997;16(8):1982–91. Pierce JV, Kumamoto CA. Variation in Candida albicans EFG1 expression enables host-dependent changes in colonizing fungal populations. MBio 2012;3(4):e00117–12. Pande K, Chen C, Noble SM. Passage through the mammalian gut triggers a phenotypic switch that promotes Candida albicans commensalism. Nat Genet 2013;45(9):1088–91. Jimenez-Lopez C, Collette JR, Brothers KM, et al. Candida albicans induces arginine biosynthetic genes in response to host-derived reactive oxygen species. Eukaryot Cell 2013;12(1):91–100. Seider K, Heyken A, Luttich A, et al. Interaction of pathogenic yeasts with phagocytes: survival, persistence and escape. Curr Opin Microbiol 2010;13(4):392–400. Lewis LE, Bain JM, Okai B, et al. Live-cell video microscopy of fungal pathogen phagocytosis. J Vis Exp 2013:pii: 50196. Wartenberg A, Linde J, Martin R, et al. Microevolution of Candida albicans in macrophages restores filamentation in a nonfilamentous mutant. PLoS Genet 2014;10(12):e1004824. Ford CB, Funt JM, Abbey D, et al. The evolution of drug resistance in clinical isolates of Candida albicans. Elife 2015;4:e00662. Hickman MA, Zeng G, Forche A, et al. The ‘obligate diploid’ Candida albicans forms mating-competent haploids. Nature 2013;494(7435):55–9. Yang F, Yan TH, Rustchenko E, et al. High-frequency genetic contents variations in clinical Candida albicans isolates. Biol Pharm Bull 2011;34(5):624–31.

91

42. Coste A, Selmecki A, Forche A, et al. Genotypic evolution of azole resistance mechanisms in sequential Candida albicans isolates. Eukaryot Cell 2007;6(10):1889–904. 43. Selmecki A, Gerami-Nejad M, Paulson C, et al. An isochromosome confers drug resistance in vivo by amplification of two genes, ERG11 and TAC1. Mol Microbiol 2008; 68(3):624–41. 44. Harrison BD, Hashemi J, Bibi M, et al. A tetraploid intermediate precedes aneuploid formation in yeasts exposed to fluconazole. PLoS Biol 2014;12(3):e1001815. 45. Selmecki A, Forche A, Berman J. Aneuploidy and isochromosome formation in drug-resistant Candida albicans. Science 2006;313(5785):367–70. 46. Selmecki AM, Dulmage K, Cowen LE, et al. Acquisition of aneuploidy provides increased fitness during the evolution of antifungal drug resistance. PLoS Genet 2009;5(10):e1000705. 47. Geber A, Hitchcock CA, Swartz JE, et al. Deletion of the Candida glabrata ERG3 and ERG11 genes: effect on cell viability, cell growth, sterol composition, and antifungal susceptibility. Antimicrob Agents Chemother 1995;39(12):2708–17. 48. Coste AT, Karababa M, Ischer F, et al. TAC1, transcriptional activator of CDR genes, is a new transcription factor involved in the regulation of Candida albicans ABC transporters CDR1 and CDR2. Eukaryot Cell 2004;3(6):1639–52. 49. Selmecki A, Bergmann S, Berman J. Comparative genome hybridization reveals widespread aneuploidy in Candida albicans laboratory strains. Mol Microbiol 2005;55(5):1553–65. 50. Chen X, Magee BB, Dawson D, et al. Chromosome 1 trisomy compromises the virulence of Candida albicans. Mol Microbiol 2004;51(2):551–65. 51. Abbey D, Hickman M, Gresham D, et al. High-resolution SNP/ CGH microarrays reveal the accumulation of loss of heterozygosity in commonly used Candida albicans Strains. G3 (Bethesda) 2011;1(7):523–30. 52. Forche A, Abbey D, Pisithkul T, et al. Stress alters rates and types of loss of heterozygosity in Candida albicans. MBio 2011;2(4):pii: e00129-11. 53. Forche A, May G, Magee PT. Demonstration of loss of heterozygosity by single-nucleotide polymorphism microarray analysis and alterations in strain morphology in Candida albicans strains during infection. Eukaryot Cell 2005;4(1):156–65. 54. Andersen MP, Nelson ZW, Hetrick ED, et al. A genetic screen for increased loss of heterozygosity in Saccharomyces cerevisiae. Genetics 2008;179(3):1179–95. 55. Coste A, Turner V, Ischer F, et al. A mutation in Tac1p, a transcription factor regulating CDR1 and CDR2, is coupled with loss of heterozygosity at chromosome 5 to mediate antifungal resistance in Candida albicans. Genetics 2006;172(4):2139–56. 56. Chang FM, Ou TY, Cheng WN, et al. Short-term exposure to fluconazole induces chromosome loss in Candida albicans: an approach to produce haploid cells. Fungal Genet Biol 2014;70:68–76. 57. Chiang DY, Getz G, Jaffe DB, et al. High-resolution mapping of copy-number alterations with massively parallel sequencing. Nat Methods 2009;6(1):99–103. 58. Peterson BK, Weber JN, Kay EH, et al. Double digest RADseq: an inexpensive method for de novo SNP discovery and genotyping in model and non-model species. PLoS One 2012;7(5):e37135. 59. Davey JW, Hohenlohe PA, Etter PD, et al. Genome-wide genetic marker discovery and genotyping using nextgeneration sequencing. Nat Rev Genet 2011;12(7):499–510.

92

| Anderson and Bennett

60. Hickman MA, Paulson C, Dudley AM, et al. Parasexual ploidy reduction drives population heterogeneity through random and transient aneuploidy in Candida albicans. Genetics 2015;200(3):781–94. 61. Catchen JM, Amores A, Hohenlohe P, et al. Stacks: building and genotyping loci de novo from short-read sequences. G3 (Bethesda), 2011;1(3):171–82. 62. Abbey DA, Funt J, Lurie-Weinberger MN, et al. YMAP: a pipeline for visualization of copy number variation and loss of heterozygosity in eukaryotic pathogens. Genome Med 2014;6(11):100. 63. Stevenson JS, Liu H. Regulation of white and opaque celltype formation in Candida albicans by Rtt109 and Hst3. Mol Microbiol 2011;81(4):1078–91. 64. Lopes da Rosa J, Boyartchuk VL, Zhu LJ, et al. Histone acetyltransferase Rtt109 is required for Candida albicans pathogenesis. Proc Natl Acad Sci USA 2010;107(4):594–9. 65. Sellam A, Askew C, Epp E, et al. Genome-wide mapping of the coactivator Ada2p yields insight into the functional roles of SAGA/ADA complex in Candida albicans. Mol Biol Cell 2009;20(9):2389–400. 66. Raman SB, Nguyen MH, Zhang Z, et al. Candida albicans SET1 encodes a histone 3 lysine 4 methyltransferase that contributes to the pathogenesis of invasive candidiasis. Mol Microbiol 2006;60(3):697–709. 67. Hasim S, Tati S, Madayiputhiya N, et al. Histone biotinylation in Candida albicans. FEMS Yeast Res 2013;13(6):529–39. 68. Berger SL. The complex language of chromatin regulation during transcription. Nature 2007;447(7143):407–12. 69. Hnisz D, Bardet AF, Nobile CJ, et al. A histone deacetylase adjusts transcription kinetics at coding sequences during Candida albicans morphogenesis. PLoS Genet 2012;8(12):e1003118. 70. Lohse MB, Johnson AD. White-opaque switching in Candida albicans. Curr Opin Microbiol 2009;12(6):650–4. 71. Soll DR. The role of phenotypic switching in the basic biology and pathogenesis of Candida albicans. J Oral Microbiol 2014;6. 72. Zordan RE, Galgoczy DJ, Johnson AD. Epigenetic properties of white-opaque switching in Candida albicans are based on a self-sustaining transcriptional feedback loop. Proc Natl Acad Sci USA 2006;103(34):12807–12. 73. Huang G, Wang H, Chou S, et al. Bistable expression of WOR1, a master regulator of white-opaque switching in Candida albicans. Proc Natl Acad Sci USA 2006;103(34):12813–18. 74. Srikantha T, Borneman AR, Daniels KJ, et al. TOS9 regulates white-opaque switching in Candida albicans. Eukaryot Cell 2006;5(10):1674–87. 75. Hernday AD, Lohse MB, Fordyce PM, et al. Structure of the transcriptional network controlling white-opaque switching in Candida albicans. Mol Microbiol 2013;90(1):22–35. 76. Wang H, Song W, Huang G, et al. Candida albicans Zcf37, a zinc finger protein, is required for stabilization of the white state. FEBS Lett 2011;585(5):797–802. 77. Lohse MB, Hernday AD, Fordyce PM, et al. Identification and characterization of a previously undescribed family of sequence-specific DNA-binding domains. Proc Natl Acad Sci USA 2013;110(19):7660–5. 78. Lassak T, Schneider E, Bussmann M, et al. Target specificity of the Candida albicans Efg1 regulator. Mol Microbiol 2011;82(3):602–18. 79. Maerkl SJ, Quake SR. A systems approach to measuring the binding energy landscapes of transcription factors. Science 2007;315(5809):233–7.

80. Fordyce PM, Gerber D, Tran D, et al. De novo identification and biophysical characterization of transcription-factor binding sites with microfluidic affinity analysis. Nat Biotechnol 2010;28(9):970–5. 81. Nobile CJ, Fox EP, Nett JE, et al. A recently evolved transcriptional network controls biofilm development in Candida albicans. Cell 2012;148(1–2):126–38. 82. Finkel JS, Mitchell AP. Genetic control of Candida albicans biofilm development. Nat Rev Microbiol 2011;9(2):109–18. 83. Park YN, Daniels KJ, Pujol C, et al. Candida albicans forms a specialized “sexual” as well as “pathogenic” biofilm. Eukaryot Cell 2013;12(8):1120–31. 84. Alby K, Bennett RJ. Interspecies pheromone signaling promotes biofilm formation and same-sex mating in Candida albicans. Proc Natl Acad Sci USA 2011;108(6):2510–15. 85. Sorrells TR, Johnson AD. Making sense of transcription networks. Cell 2015;161(4):714–23. 86. Perez JC, Fordyce PM, Lohse MB, et al. How duplicated transcription regulators can diversify to govern the expression of nonoverlapping sets of genes. Genes Dev 2014;28(12): 1272–7. 87. Tsong AE, Tuch BB, Li H, et al. Evolution of alternative transcriptional circuits with identical logic. Nature 2006;443(7110):415–20. 88. Tsong AE, Miller MG, Raisner RM, et al. Evolution of a combinatorial transcriptional circuit: a case study in yeasts. Cell 2003;115(4):389–99. 89. Baker CR, Booth LN, Sorrells TR, et al. Protein modularity, cooperative binding, and hybrid regulatory states underlie transcriptional network diversification. Cell 2012;151(1):80–95. 90. Lavoie H, Hogues H, Mallick J, et al. Evolutionary tinkering with conserved components of a transcriptional regulatory network. PLoS Biol 2010;8(3):e1000329. 91. Lane S, Birse C, Zhou S, et al. DNA array studies demonstrate convergent regulation of virulence factors by Cph1, Cph2, and Efg1 in Candida albicans. J Biol Chem 2001;276(52): 48988–96. 92. Murad AM, d’Enfert C, Gaillardin C, et al. Transcript profiling in Candida albicans reveals new cellular functions for the transcriptional repressors CaTup1, CaMig1 and CaNrg1. Mol Microbiol 2001;42(4):981–93. 93. De Backer MD, Ilyina T, Ma XJ, et al. Genomic profiling of the response of Candida albicans to itraconazole treatment using a DNA microarray. Antimicrob Agents Chemother 2001;45(6):1660–70. 94. Cowen LE, Nantel A, Whiteway MS, et al. Population genomics of drug resistance in Candida albicans. Proc Natl Acad Sci USA 2002;99(14):9284–9. 95. Nantel A, Dignard D, Bachewich C, et al. Transcription profiling of Candida albicans cells undergoing the yeastto-hyphal transition. Mol Biol Cell 2002;13(10):3452–65. 96. Sellam A, Hogues H, Askew C, et al. Experimental annotation of the human pathogen Candida albicans coding and noncoding transcribed regions using high-resolution tiling arrays. Genome Biol 2010;11(7):R71. 97. Bruno VM, Wang Z, Marjani SL, et al. Comprehensive annotation of the transcriptome of the human fungal pathogen Candida albicans using RNA-seq. Genome Res 2010;20(10):1451–8. 98. Tuch BB, Mitrovich QM, Homann OR, et al. The transcriptomes of two heritable cell types illuminate the circuit governing their differentiation. PLoS Genet 2010;6(8):e1001070. 99. Wilson D, Thewes S, Zakikhany K, et al. Identifying infection-associated genes of Candida albicans in the postgenomic era. FEMS Yeast Res 2009;9(5):688–700.

Bringing functional genomics to Candida albicans |

100. Walker LA, Maccallum DM, Bertram G, et al. Genome-wide analysis of Candida albicans gene expression patterns during infection of the mammalian kidney. Fungal Genet Biol 2009;46(2):210–19. 101. Uppuluri P, Chaffin WL. Defining Candida albicans stationary phase by cellular and DNA replication, gene expression and regulation. Mol Microbiol 2007;64(6):1572–86. 102. Cairns T, Minuzzi F, Bignell E. The host-infecting fungal transcriptome. FEMS Microbiol Lett 2010;307(1):1–11. 103. Kadosh D, Johnson AD. Induction of the Candida albicans filamentous growth program by relief of transcriptional repression: a genome-wide analysis. Mol Biol Cell 2005;16(6):2903–12. 104. Tao L, Du H, Guan G, et al. Discovery of a “whitegray-opaque” tristable phenotypic switching system in Candida albicans: roles of non-genetic diversity in host adaptation. PLoS Biol 2014;12(4):e1001830. 105. Traven A, Janicke A, Harrison P, et al. Transcriptional profiling of a yeast colony provides new insight into the heterogeneity of multicellular fungal communities. PLoS One 2012;7(9):e46243. 106. Cleary IA, Lazzell AL, Monteagudo C, et al. BRG1 and NRG1 form a novel feedback circuit regulating Candida albicans hypha formation and virulence. Mol Microbiol 2012;85(3):557–73. 107. Muzzey D, Sherlock G, Weissman JS. Extensive and coordinated control of allele-specific expression by both transcription and translation in Candida albicans. Genome Res 2014;24(6):963–73. 108. Muzzey D, Schwartz K, Weissman JS, et al. Assembly of a phased diploid Candida albicans genome facilitates allele-specific measurements and provides a simple model for repeat and indel structure. Genome Biol 2013;14(9):R97. 109. Fradin C, De Groot P, MacCallum D, et al. Granulocytes govern the transcriptional response, morphology and proliferation of Candida albicans in human blood. Mol Microbiol 2005;56(2):397–415. 110. Lorenz MC, Bender JA, Fink GR. Transcriptional response of Candida albicans upon internalization by macrophages. Eukaryot Cell 2004;3(5):1076–87. 111. Barelle CJ, Richard ML, Gaillardin C, et al. Candida albicans VAC8 is required for vacuolar inheritance and normal hyphal branching. Eukaryot Cell 2006;5(2):359–67. 112. Enjalbert B, MacCallum DM, Odds FC, et al. Niche-specific activation of the oxidative stress response by the pathogenic fungus Candida albicans. Infect Immun 2007;75(5):2143–51. 113. Spiering MJ, Moran GP, Chauvel M, et al. Comparative transcript profiling of Candida albicans and Candida dubliniensis identifies SFL2, a C. albicans gene required for virulence in a reconstituted epithelial infection model. Eukaryot Cell 2010;9(2):251–65. 114. Zakikhany K, Naglik JR, Schmidt-Westhausen A, et al. In vivo transcript profiling of Candida albicans identifies a gene essential for interepithelial dissemination. Cell Microbiol 2007;9(12):2938–54. 115. Bruno VM, Shetty AC, Yano J, et al. Transcriptomic analysis of vulvovaginal candidiasis identifies a role for the NLRP3 inflammasome. MBio 2015;6(2):e00182–15. 116. Naglik JR, Moyes D, Makwana J, et al. Quantitative expression of the Candida albicans secreted aspartyl proteinase gene family in human oral and vaginal candidiasis. Microbiology 2008;154(Pt 11):3266–80. 117. Liu Y, Shetty AC, Schwartz JA, et al. New signaling pathways govern the host response to C. albicans infection in various niches. Genome Res 2015;25(5):679–89.

93

118. Andes D, Lepak A, Pitula A, et al. A simple approach for estimating gene expression in Candida albicans directly from a systemic infection site. J Infect Dis 2005;192(5):893–900. 119. Xu W, Solis NV, Ehrlich RL, et al. Activation and alliance of regulatory pathways in C. albicans during mammalian infection. PLoS Biol 2015;13(2):e1002076. 120. Kulkarni MM. Digital multiplexed gene expression analysis using the NanoString nCounter system. Curr Protoc Mol Biol 2011;Chapter 25:Unit25B 10. 121. Chen YY, Chao CC, Liu FC, et al. Dynamic transcript profiling of Candida albicans infection in zebrafish: a pathogen-host interaction study. PLoS One 2013;8(9):e72483. 122. Rosenbach A, Dignard D, Pierce JV, et al. Adaptations of Candida albicans for growth in the mammalian intestinal tract. Eukaryot Cell 2010;9(7):1075–86. 123. Pierce JV, Dignard D, Whiteway M, et al. Normal adaptation of Candida albicans to the murine gastrointestinal tract requires Efg1p-dependent regulation of metabolic and host defense genes. Eukaryot Cell 2013;12(1):37–49. 124. Skrzypek MS, Arnaud MB, Costanzo MC, et al. New tools at the Candida Genome Database: biochemical pathways and full-text literature search. Nucleic Acids Res 2010;38:D428–32. 125. Sellam A, van het Hoog M, Tebbji F, et al. Modeling the transcriptional regulatory network that controls the early hypoxic response in Candida albicans. Eukaryot Cell 2014;13(5):675–90. 126. Dhamgaye S, Bernard M, Lelandais G, et al. RNA sequencing revealed novel actors of the acquisition of drug resistance in Candida albicans. BMC Genomics 2012;13:396. 127. Martin R, Albrecht-Eckardt D, Brunke S, et al. A core filamentation response network in Candida albicans is restricted to eight genes. PLoS One 2013;8(3):e58613. 128. Novershtern N, Subramanian A, Lawton LN, et al. Densely interconnected transcriptional circuits control cell states in human hematopoiesis. Cell 2011;144(2):296–309. 129. Bolouri H. Modeling genomic regulatory networks with big data. Trends Genet 2014;30(5):182–91. 130. Ellwanger DC, Leonhardt JF, Mewes HW. Large-scale modeling of condition-specific gene regulatory networks by information integration and inference. Nucleic Acids Res 2014;42(21):doi: 10.1093/nar/gku916. 131. White TC, Andrews LE, Maltby D, et al. The “universal” leucine codon CTG in the secreted aspartyl proteinase 1 (SAP1) gene of Candida albicans encodes a serine in vivo. J Bacteriol 1995;177(10):2953–5. 132. Bennett RJ, Johnson AD. Mating in Candida albicans and the search for a sexual cycle. Annu Rev Microbiol 2005;59:233–55. 133. Enloe B, Diamond A, Mitchell AP. A single transformation gene function test in diploid Candida albicans. J Bacteriol 2000;182(20):5730–6. 134. Davis DA, Bruno VM, Loza L, et al. Candida albicans Mds3p, a conserved regulator of pH responses and virulence identified through insertional mutagenesis. Genetics 2002;162(4):1573–81. 135. Spreghini E, Davis DA, Subaran R, et al. Roles of Candida albicans Dfg5p and Dcw1p cell surface proteins in growth and hypha formation. Eukaryot Cell 2003;2(4):746–55. 136. Mukherjee PK, Mohamed S, Chandra J, et al. Alcohol dehydrogenase restricts the ability of the pathogen Candida albicans to form a biofilm on catheter surfaces through an ethanol-based mechanism. Infect Immun 2006;74(7):3804–16.

94

| Anderson and Bennett

137. Xu W, Smith FJ, Jr, Subaran R, et al. Multivesicular body-ESCRT components function in pH response regulation in Saccharomyces cerevisiae and Candida albicans. Mol Biol Cell 2004;15(12):5528–37. 138. Nobile CJ, Mitchell AP. Regulation of cell-surface genes and biofilm formation by the C. albicans transcription factor Bcr1p. Curr Biol 2005;15(12):1150–5. 139. Blankenship JR, Fanning S, Hamaker JJ, et al. An extensive circuitry for cell wall regulation in Candida albicans. PLoS Pathog 2010;6(2):e1000752. 140. Bruno VM, Kalachikov S, Subaran R, et al. Control of the C. albicans cell wall damage response by transcriptional regulator Cas5. PLoS Pathog 2006;2(3):e21. 141. Uhl MA, Biery M, Craig N, et al. Haploinsufficiency-based large-scale forward genetic analysis of filamentous growth in the diploid human fungal pathogen C. albicans. EMBO J 2003;22(11):2668–78. 142. Oh J, Fung E, Schlecht U, et al. Gene annotation and drug target discovery in Candida albicans with a tagged transposon mutant collection. PLoS Pathog 2010;6(10):e1001140. 143. Roemer T, Jiang B, Davison J, et al. Large-scale essential gene identification in Candida albicans and applications to antifungal drug discovery. Mol Microbiol 2003;50(1):167–81. 144. Xu D, Sillaots S, Davison J, et al. Chemical genetic profiling and characterization of small-molecule compounds that affect the biosynthesis of unsaturated fatty acids in Candida albicans. J Biol Chem 2009;284(29):19754–64. 145. Tebbji F, Chen Y, Richard Albert J, et al. A functional portrait of Med7 and the mediator complex in Candida albicans. PLoS Genet 2014;10(11):e1004770. 146. Noble SM, French S, Kohn LA, et al. Systematic screens of a Candida albicans homozygous deletion library decouple morphogenetic switching and pathogenicity. Nat Genet 2010;42(7):590–8. 147. Homann OR, Dea J, Noble SM, et al. A phenotypic profile of the Candida albicans regulatory network. PLoS Genet 2009;5(12):e1000783. 148. Perez JC, Kumamoto CA, Johnson AD. Candida albicans commensalism and pathogenicity are intertwined traits directed by a tightly knit transcriptional regulatory circuit. PLoS Biol 2013;11(3):e1001510. 149. Cabral V, Chauvel M, Firon A, et al. Modular gene overexpression strategies for Candida albicans. Methods Mol Biol 2012;845:227–44. 150. Jin R, Dobry CJ, McCown PJ, et al. Large-scale analysis of yeast filamentous growth by systematic gene disruption and overexpression. Mol Biol Cell 2008;19(1):284–96.

151. Care RS, Trevethick J, Binley KM, et al. The MET3 promoter: a new tool for Candida albicans molecular genetics. Mol Microbiol 1999;34(4):792–8. 152. Backen AC, Broadbent ID, Fetherston RW, et al. Evaluation of the CaMAL2 promoter for regulated expression of genes in Candida albicans. Yeast 2000;16(12):1121–9. 153. Nakayama H, Mio T, Nagahashi S, et al. Tetracyclineregulatable system to tightly control gene expression in the pathogenic fungus Candida albicans. Infect Immun 2000;68(12):6712–19. 154. Park YN, Morschhauser J. Tetracycline-inducible gene expression and gene deletion in Candida albicans. Eukaryot Cell 2005;4(8):1328–42. 155. Sahni N, Yi S, Daniels KJ, et al. Tec1 mediates the pheromone response of the white phenotype of Candida albicans: insights into the evolution of new signal transduction pathways. PLoS Biol 2010;8(5):e1000363. 156. Du H, Guan G, Xie J, et al. Roles of Candida albicans Gat2, a GATA-type zinc finger transcription factor, in biofilm formation, filamentous growth and virulence. PLoS One 2012;7(1):e29707. 157. Vashee S, Kodadek T. The activation domain of GAL4 protein mediates cooperative promoter binding with general transcription factors in vivo. Proc Natl Acad Sci USA 1995;92(23):10683–7. 158. Schillig R, Morschhauser J. Analysis of a fungus-specific transcription factor family, the Candida albicans zinc cluster proteins, by artificial activation. Mol Microbiol 2013;89(5):1003–17. 159. Chauvel M, Nesseir A, Cabral V, et al. A versatile overexpression strategy in the pathogenic yeast Candida albicans: identification of regulators of morphogenesis and fitness. PLoS One 2012;7(9):e45912. 160. Leuker CE, Sonneborn A, Delbruck S, et al. Sequence and promoter regulation of the PCK1 gene encoding phosphoenolpyruvate carboxykinase of the fungal pathogen Candida albicans. Gene 1997;192(2):235–40. 161. Vyas VK, Barrasa MI, Fink GR. A CRISPR system permits genetic engineering of essential genes and gene families. Sci Adv 2015;1(3):e1500248. 162. Landt SG, Marinov GK, Kundaje A, et al. ChIP-seq guidelines and practices of the ENCODE and modENCODE consortia. Genome Res 2012;22(9):1813–31. 163. Teytelman L, Thurtle DM, Rine J, et al. Highly expressed loci are vulnerable to misleading ChIP localization of multiple unrelated proteins. Proc Natl Acad Sci USA 2013;110(46):18602–7. 164. Sims D, Sudbery I, Ilott NE, et al. Sequencing depth and coverage: key considerations in genomic analyses. Nat Rev Genet 2014;15(2):121–32.

Budding off: bringing functional genomics to Candida albicans.

Candida species are the most prevalent human fungal pathogens, with Candida albicans being the most clinically relevant species. Candida albicans resi...
NAN Sizes 0 Downloads 9 Views