Next Article in Journal
The Use of Raman Spectroscopy to Monitor Metabolic Changes in Stressed Metschnikowia sp. Yeasts
Previous Article in Journal
Citrobacter tructae sp. nov. Isolated from Kidney of Diseased Rainbow Trout (Oncorhynchus mykiss)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Systematic Analysis of Functionally Related Gene Clusters in the Opportunistic Pathogen, Candida albicans

Department of Biology, William Paterson University, Wayne, NJ 07470, USA
*
Author to whom correspondence should be addressed.
Submission received: 29 December 2020 / Accepted: 20 January 2021 / Published: 28 January 2021
(This article belongs to the Special Issue Genome Evolution in Fungal Pathogens)

Abstract

:
The proper balance of gene expression is essential for cellular health, organismal development, and maintaining homeostasis. In response to complex internal and external signals, the cell needs to modulate gene expression to maintain proteostasis and establish cellular identity within its niche. On a genome level, single-celled prokaryotic microbes display clustering of co-expressed genes that are regulated as a polycistronic RNA. This phenomenon is largely absent from eukaryotic microbes, although there is extensive clustering of co-expressed genes as functional pairs spread throughout the genome in Saccharomyces cerevisiae. While initial analysis demonstrated conservation of clustering in divergent fungal lineages, a comprehensive analysis has yet to be performed. Here we report on the prevalence, conservation, and significance of the functional clustering of co-regulated genes within the opportunistic human pathogen, Candida albicans. Our analysis reveals that there is extensive clustering within this organism—although the identity of the gene pairs is unique compared with those found in S. cerevisiae—indicating that this genomic arrangement evolved after these microbes diverged evolutionarily, rather than being the result of an ancestral arrangement. We report a clustered arrangement in gene families that participate in diverse molecular functions and are not the result of a divergent orientation with a shared promoter. This arrangement coordinates the transcription of the clustered genes to their neighboring genes, with the clusters congregating to genomic loci that are conducive to transcriptional regulation at a distance.

1. Introduction

Many Candida species are opportunistic pathogens, with cells undergoing profound transcriptional changes during the transition into a pathogenic form. These pathogens are a major cause of morbidity and mortality across the world and represent a significant public health threat [1,2]. When pathogenic, Candida species can manifest as: candidemia, in the formation of biofilms, urinary tract infections, vulvovaginal infections, thrush, and associated invasive candidiasis [3,4,5,6,7]. There are a finite number of treatments available, including azoles, flucytosine and echinocandin drugs [8]. Unfortunately, multidrug resistance to available options is on the rise and driven in part by genomic instability and a mutator phenotype [8,9,10]. Further complicating the situation is the emergence of novel, more pathogenic species, such as C. auris, highlighting the need for a more complete understanding of the biology of these species. Of particular importance is the genomic organization, transcriptional regulation, and regulatory mechanisms associated with the changes in the lifecycle of Candida strains—such as during the switch to infection and pathogenesis [11,12,13,14].
Proper gene regulation is essential to all organisms to establish cellular identity and maintain homeostasis [15,16,17,18,19]. All single-celled eukaryotic organisms, including Candida albicans, deal with the complexities of this response with multiple levels of gene expression regulation [20,21,22,23,24,25]. Multiple pathways recognize signals—both internal and external—and respond, ultimately converging to alter the transcriptome of the cell, triggering distinct gene expression programs and specific transcriptional signatures [26,27]. These changes are regulated at multiple levels by myriad mechanisms, ensuring adaptation to the environmental niche that the cell occupies [28,29,30]. In single-celled organisms, including yeasts, these changes are rapid and can result in alterations to the expression of thousands of genes simultaneously to ensure survival [28,31].
The regulatory circuitry that is required to survive in the host depends, in part, on chromatin remodeling and the production of chaperone proteins [32]. Biofilm formation depends on the activity of several transcription factors that target over one thousand genes [33,34]. An additional layer of transcriptional regulation is through genomic organization. The spatial positioning of functionally related, co-expressed gene families as clusters throughout the genome is a central organizing principal in the related Ascomycete, Saccharomyces cerevisiae [35,36]. This arrangement—the functional clustering of co-expressed genes that participate in a shared molecular process—helps to minimize the effects of stochastic gene expression and noise [36,37,38]. There are significant levels of genomic clustering observed in the ribosomal protein (RP) and rRNA and ribosome biosynthesis (RRB) gene families in C. albicans, and while the absolute levels of clustering present are comparable to those in S. cerevisiae, the identity of the clusters differs [35,39,40]. At present, a systematic analysis of the prevalence and significance of genomic organization on transcriptional coregulation has yet to be performed in C. albicans.
In this work, we systematically characterize the genomic organization of functionally related, co-expressed gene families into clusters throughout the genome of the fungal pathogen, C. albicans. We report that the genomic distribution of gene families into clusters is widespread in C. albicans and that this arrangement is extensively conserved within related Candida species. The co-localization of genes as functional clusters frequently results in tighter transcriptional co-expression for the clustered genes. Altogether, our work suggests that this organization represents an integral level of organization to facilitate and coordinate the co-expression of functionally related genes.

2. Materials and Methods

2.1. The Identification and Significance of Functional Clustering in Candida albicans

The prevalence and statistical probability of genomic arrangement was determined by accessing the membership of each gene family annotated by the gene ontology (GO) molecular process designations from the Candida Genome Database [39,40,41]. The global, genomic distribution for the thirty-eight gene families was subsequently determined by manual curation throughout the Candida haploid genome, as previously described [35]. Once the family membership, locations, and orientations were mapped, the significance of this genomic arrangement was determined by calculating the binomial probability for the arrangement as previously described [42].

2.2. The Conservation of the Functional Clusters across Divergent Candida Lineages

The conservation of the observed genomic distribution was determined by manual curation from the Candida gene order browser (CGOB) [43,44]. The conservation of syntenic relationships identified in C. albicans was explored in the following Candida and yeast strains: WO1, C. dubliniensis, C. tropicalis, C. parapsilosis, C. metapsilosis, C. orthopsilosis, L. elongisporus, D. hansenii, S. stipitis, C. tenuis, S. passalidarum, M. guilliermondii, C. lusitaniae, C. auris, and S. cerevisiae. Gene clusters were queried as pairings (to account for the partial conservation of larger clusters) and the relationships were defined as conserved if the syntenic context was identical within a divergent Candida genome. Heat maps were generated to facilitate visualization in a binary fashion, with each grouping defined as conserved (yellow) or not conserved (black). Genes were arranged by family membership and then by the extent of conservation within each family, omitting ‘Cell budding’, ‘Cell wall organization’, ‘Conjugation’, ‘Protein folding’, and the ‘Transposition’ gene families, as there were no identified functional clusters in Candida.

2.3. Calculation of the Transcriptional Relationship within Gene Families with the Pearson’s Correlation Coefficient

The average pairwise Pearson’s correlation coefficient (PCC) from gene expression data was calculated as previously described [42]. The PCC scores for the unpaired genes were calculated from the average of every possible pairing partner for every possible gene. P values were determined by bootstrapping with replacement by taking at least 10,000 random groupings of genes (the same size as the paired subset) and determining the average PCC score for that grouping.

2.4. Calculation of the Transcriptional Relationship within a Genomic Neighborhood Using the Spearman’s Correlation Coefficient

Determination of transcriptional coregulation at a distance was calculated by the Spearman’s correlation coefficient (SCC). The Spearman’s correlation coefficient was determined as described previously [45]. Calculations were performed using Python and the Pandas library. Values were plotted as a function of distance, using the genomic distance between transcriptional start sites for two genes.

2.5. Microarray and RNA-Sequencing Data Sets Used for Analysis

The datasets analyzed for gene expression analysis were accessed and downloaded from the Gene Expression Omnibus in the National Library of Medicine (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/geo/). Conditions that induced a broad range of cellular transcriptional responses for analysis were selected for analysis. The microarray and RNA-sequencing gene expression datasets utilized were: the transcriptional changes in response to carbon shift from glucose to xylose (GSE50476), response to the anti-fungal thiosemicarbazone, NSC319726, (GSE106486), response to the sterol alkaloid tomatidine (GSE96965), and the response to mucin (GSE149196) [46,47,48].

3. Results

3.1. Functionally Related Genes Cluster throughout the Candida albicans Genome

The prevalence of the functional clustering in Candida albicans was determined for the thirty-eight gene families annotated by the Gene Ontology Consortium as having a shared Molecular Process [39]. The individual members of each family were accessed and the genomic location of every member was curated utilizing the Candida Genome Database to determine the occurrence of clusters within each family (Table 1) [41]. There were 38 gene families in our initial analysis, with the individual gene families varying in size from as small as four genes, as in the ‘Transposition’ gene family, up to 1500 genes, as in the ‘Regulation of biological process’ family. The broad scope of several of the larger gene families, including ‘Regulation of biological process’ and ‘Organelle organization’ (1053 members), served as controls—these families are non-specific umbrella terms that include multiple members from more specific families—and thus would not be expected to cluster to a significant degree. The composition of these groups in particular consist of members that are cross-listed under other, more specific ontology headings.
For completion of this analysis, we did not exclude any members throughout this phase of the project. A significant fraction of the queried gene families—twenty out of the thirty-eight families—exhibit a non-random distribution throughout the genome (utilizing a threshold for significance of p < 0.05). Eight families exhibit a highly significant distribution (p < 0.001): ‘Biofilm formation’, ‘Carbohydrate metabolic process’, ‘Cell adhesion’, ‘Cell cycle’, ‘Cytoskeleton organization’, ‘Generation of precursor metabolites and energy’, ‘Translation’, and ‘Vitamin metabolic process’ (Table 1). As expected, the largest gene families, that represented umbrella terms and are composed of genes from a variety of more specific molecular functional groups exhibited a random distribution using our statistical cutoff.

3.2. Functionally Clustered Genes Exhibit Tighter Transcriptional Coregulation within Their Gene Family

In order to determine the effect of clustering on the transcription within a co-regulated gene family, the Pearson’s correlation coefficient (PCC) was calculated throughout gene expression changes induced by a variety of growth and lifestyle conditions (complete conditions and datasets are listed in the Materials and Methods section). Microarray and RNA-sequencing (RNA-seq) gene expression datasets were accessed and the transcription profiles for each of the gene families that displayed a clustered organization were extracted (the five gene families that contained zero clusters among their membership were not included in these analyses).
The average pairwise PCC values for the entire gene family, the pairwise PCC values for the clustered members of the gene family, and the pairwise PCC values for the unclustered, or singleton, members of each gene family were calculated (Table 2). Our analysis shows that across all conditions, all gene families studied have a positive correlation, although there is significant variation between families—with the ‘Ribosome biogenesis’, ‘Nucleus organization’, ‘Translation’, and ‘RNA metabolic processes’ displaying the tightest transcriptional coordination. Using a cutoff of a difference in transcriptional coordination of ≥0.1, we identified six gene families where the clustered genes are more tightly co-expressed together and another ten families where there was a slight increase relative to the singleton family members. Surprisingly, eight families have a more significant correlation for the singletons rather than the clustered set of genes. One gene family, ‘Pseudohyphal growth’, demonstrated an anti-correlation that is characteristic of mutually exclusive gene expression, potentially by transcriptional interference.

3.3. Clustered Genes Do Not Exhibit a Bias for a Divergent Orientation

There are three potential orientations for clustered genes, which can be found in a divergent (←→), tandem (→→ or ←←), or convergent orientation (→←) with respect to their transcription start sites. Of the three possibilities, a divergent orientation provides a facile model for the coregulation of a gene cluster via the function of a shared bi-directional promoter. In order to ascertain if the gene clusters were biased towards this particular mechanism, the relative orientations for every functional cluster were determined and annotated (Table 3).
For this analysis, clusters that appeared under multiple ontological headings were eliminated, keeping only the most specific assigned molecular function. This reduced the number of genes in this analysis significantly—from a total of 2945 total clustered members to 1452 unique clusters spanning all 33 ontology categories. There is no observable bias towards a divergent orientation, with an approximately even split between divergent (←→), where there are 389 pairings found in this orientation, and convergent (→←), where there are 343 pairings found in this orientations. The only observable bias is towards a tandem (→→ or ←←) orientation, which is overrepresented relative to the other two possibilities with a total of 720 pairings in this orientation.

3.4. Clustered Grouping Are Highly Conserved among Closely Related Candida Species and Deteriorate with Greater Evolutionary Distance

To characterize the relationship between clustered genes within this particular clade, the conservation of the identity of clustered genes was determined across extant Candida lineages through the budding yeast, Saccharomyces cerevisiae. The unique set of clustered genes was individually queried from Candida albicans as a reference point and were assigned a binary score as either conserved or not conserved for this analysis. Analysis is performed in C. dubliniensis, C. tropicalis, C. parapsilosis, C. metapsilosis, C. orthopsilosis, L. elongisporus, D. hansenii, S. stipitis, C. tenuis, S. passalidarum, M. guilliermondii, C. lusitaniae, C. auris, and S. cerevisiae, and is visualized as a heat map (Figure 1). For simplicity, individual clusters are grouped by family and then by the extent of conservation within the family. The identity of clusters was highly conserved within the closely related Candida species, consistent with previous analysis performed across Saccharomyces lineages [42]. The conservation of clustered gene identity decreases as the evolutionary distance between Candida albicans and the fungal lineage for comparison increased, consistent with previous observations. The conservation eroded significantly when compared to the distantly related budding yeast, S. cerevisiae, indicating that the majority of clustered relationships most likely arose after the evolutionary divergence between these species.

3.5. Functionally Clustered Gene Sets Display a Higher Transcriptional Coregulation than Would Be Expected by Chance Groupings of Functionally Related Genes

Previous studies in yeasts have characterized the set of clustered genes as distinct from the many possible clustered combinations that could have arisen by evolutionary chance [42]. In order to test the transcriptional effect of clustering to that of all the genes that could have been grouped together, a bootstrapping analysis was performed with the actual grouping’s PCC denoted with an asterisk (Figure 2). By extending this type of analysis to the Candida albicans clustered genes set, we report a significant number of grouping combinations that have a tighter transcription profile than the actual paired set—even in some of the grouping classifications that contained the highest PCC values, such as ‘Ribosome biogenesis’. There is a correlation identified within the non-random clustered sets, where the ‘Biofilm formation’, ‘Carbohydrate metabolic process’, ‘Cell adhesion’, ‘Generation of precursor metabolites and energy’, ‘Lipid metabolic process’, ‘Protein catabolic process’, ‘RNA metabolic process’, and ‘Vitamin metabolic process’ all exhibit a higher than expected PCC by chance alone. The control sets (non-significant genomic distribution) that include the ‘Regulation of biological process’, ‘Response to drug’, and ‘Response to stress’ do not exhibit the same pattern. It is interesting that in all cases, the actual clustered sets all exhibit a higher PCC value than the least clustered members of the sets, in many cases bordering on the threshold for statistical significance. This is consistent with the role that cis- and trans-factors play in coordinating the transcription of each functional set of genes, with spatial positioning potentially reinforcing this process during the stress conditions explored in our analysis.

3.6. Functionally Related Genes Congregate to Transcriptionally Permissive Regions throughout the Genome

Functionally related genes are clustered into genomic loci that exhibit a stronger influence on the expression of their neighboring genes in certain yeasts [45]. In order to determine if the Candida albicans functional clusters are found in loci with similar properties, the Spearman’s correlation coefficient (SCC) was calculated between each gene and their ten closest neighboring genes, which was then plotted as a function of genomic distance. A line of best fit was plotted over the data to facilitate interpretation. Our analysis focused on a global analysis, which was compared to an analysis of the clustered set of genes (Figure 3A,B). Consistent with observations seen across eukaryotic lineages, there is a global transcriptional similarity (positive SCC value) between all genes that decays as a function of distance (Figure 3A). The transcriptional similarity at loci where the clustered set of genes congregates displays a similar pattern, however at close genomic distances, there is a larger SCC value (Figure 3B). For ease of comparison, the two plots were superimposed on the same graph, scaled to view the global SCC and then zoomed into the proximate region alone (Figure 3C,D, respectively). This comparison confirms that the clustered loci are found in regions that exhibit a stronger transcriptional similarity as measured by the SCC than a typical genomic locus. The rate of decay is similar and approaches the same basal level that is observed across the conditions analyzed. One surprising observation was that the decay does not approach zero in our analysis; it approaches a value of approximately 0.3. This is most likely due to the nature of the conditions analyzed, which typically invoke a stress (or stress-like) response in this organism. As this is the case we believe that the positive value rather than a zero value at large genomic distances is an artifact of the data sets analyzed, rather than indicative of a global positive SCC value irrespective of genomic distance in this organism, and would not be predicted to be observed in cycling and unstressed cells populations.

4. Discussion

The clustering of functionally related gene families in the budding yeast, S. cerevisiae, is an organizational feature of the gene sets that contribute to ‘adjacent gene coregulation’ [49,50]. While initially recognized in the ‘Ribosomal protein’ and the ‘rRNA and ribosome biogenesis’ gene families, clustering has subsequently been identified in numerous gene families in S. cerevisiae [35,51]. Here, our analysis expands the characterization of the functional clustering of co-regulated genes to Candida albicans and related Candida lineages. We report that there is extensive gene clustering observed in functionally related, co-expressed gene families in C. albicans. This is conserved, at least in part, throughout related Candida lineages. The extent of the conservation of clustering correlates with evolutionary distance between species compared.
One surprising observation is that even when there are comparable levels of clustering within a co-regulated gene family, the identity of the clustered members differ between divergent fungi. Taking the ‘Ribosomal protein’, ‘rRNA and ribosome biogenesis’, and ‘Vitamin metabolic process’ gene families as examples, there are comparable levels of clustering observed in all three of these families in both C. albicans and in S. cerevisiae. The individual membership of each family is not conserved, as the gene members that comprise the functional clusters differ in each species. While this is consistent with previous observations, this present work expands upon this greatly, supporting a model where clustering is occurring in different species via mechanisms that converge at this arrangement [35,49]. This overlap of clustered genes is minimal when compared to S. cerevisiae—consistent with previous findings taken from that species—indicating that the majority of the clustered genes within each gene family arose independently within these two lineages [35].
When clustered together, the genes exhibit a coupled transcriptional correlation in many cases. These most likely result in a fitness advantage via the minimization of the stochastic influences on transcription. The mechanism that links transcription within clusters has yet to be decisively elucidated, however there is evidence supporting the role of chromatin remodelers mediating changes at the local level to modulated transcription within a genomic neighborhood [35]. While genes that are found in a divergent orientation oftentimes have bidirectional promoters, as seen in the histone protein coding genes, this does not appear to be a major factor in the regulation of clusters in C. albicans [52,53]. The fact that there is a disproportionate incidence of gene clusters found in a tandem orientation is surprising—there are incidences of tandem oriented pairs exhibiting mutually exclusive transcription patterns, as in SER3 transcription and heat shock protein gene family [49,54,55]. This could explain the number of cases where there was a less positive PCC for clustered genes. Positive transcriptional regulation in this manner could be the result of bicistronic transcripts, which are under-characterized in yeasts. It would be fascinating to see if this mechanism is utilized to a significant degree in Candida albicans.
The benefit of clustering co-expressed genes together is two-fold. This eliminates the potential incidence of toxic intermediate compounds within biosynthetic pathways and it facilitates stoichiometric levels of proteins that are necessary in specific complexes—both of which confer a survival benefit and could serve as the selection mechanism for this organization [37,38]. We postulate that the drivers of this clustering arrangement are likely a combination of these two mechanisms. The ability to maintain proteostasis is vital, as orphan proteins that are not incorporated into a complex can induce proteotoxic stress and decrease viability, as seen in the RP family [38]. A clustered arrangement may help to buffer the stochastic noise inherent in gene expression and link the expression of genes to minimize this effect. This would explain why there are comparable levels of clustering in certain families, albeit with the different composite memberships that we observe. If two (or more) genes were clustered in a manner that provides a protective benefit by limiting toxic intermediate compounds within a pathway, we would potentially expect those genes to be clustered across a long evolutionary distance. The galactose and biotin biosynthetic pathways are two examples of these groupings, and indeed they are clustered together in such a fashion [36]. While beyond the scope of this work, it would be fascinating to see if the clusters that are conserved over longer distances do have this relationship in Candida species—such an analysis could potentially identify toxic factors during infectivity, and there are oftentimes detoxification mechanisms clustered with the toxin synthesizing enzymes to protect the cell [36].

5. Conclusions

Though it has been observed that genome-wide global transcriptional similarities have been observed in yeasts and other eukaryotes for a long time, the functional benefit to eukaryotes is still in the process of being fully understood [54,55]. The physical arrangement of genes along the chromosome coordinates transcription of functionally related genes at many loci. This allows for tighter, more efficient transcription of genes whose functions participate in a shared molecular process by the process of adjacent gene coregulation [49]. Alternatively, this configuration can facilitate mutually exclusive expression of select pairings via transcriptional interference. Candida albicans and related Candida lineages exhibit extensive genomic organization in this manner and this may play a significant role in organismal adaptation, survival, and pathogenesis.

Author Contributions

Methodology, software, and analysis, S.A.; methodology, software, and analysis, R.E.; methodology, analysis, data curation, K.S.; analysis and data curation, J.M., B.P., D.K., and U.S.; conceptualization, supervision, analysis, data curation, writing and revising the manuscript, and project administration J.T.A. All authors have read and agreed to the published version of the manuscript.

Funding

NSF GS-LSAMP provided undergraduate support to KS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Complete information on the Datasets utilized for analysis is provided in the Methods section of this manuscript.

Acknowledgments

The authors would like to acknowledge Jonathan J. Foley, IV. For his invaluable assistance and advice during this project. J.T.A would like to acknowledge support from the Provost of William Paterson University for Assigned Release Time that facilitated completion of this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pappas, P.G.; Kauffman, C.A.; Andes, D.R.; Clancy, C.J.; Marr, K.A.; Ostrosky-Zeichner, L.; Reboli, A.C.; Schuster, M.G.; Vazquez, J.A.; Walsh, T.J. Clinical practice guideline for the management of candidiasis: 2016 update by the Infectious Diseases Society of America. Clin. Infect. Dis. 2016, 62, e1–e50. [Google Scholar] [CrossRef]
  2. De Oliveira Santos, G.C.; Vasconcelos, C.C.; Lopes, A.J.; de Sousa Cartágenes, M.d.S.; do Nascimento, F.R.; Ramos, R.M.; Pires, E.R.; de Andrade, M.S.; Rocha, F.M.; de Andrade Monteiro, C. Candida infections and therapeutic strategies: Mechanisms of action for traditional and alternative agents. Front. Microbiol. 2018, 9, 1351. [Google Scholar] [CrossRef] [PubMed]
  3. Nett, J.E.; Andes, D.R. Contributions of the biofilm matrix to Candida pathogenesis. J. Fungi 2020, 6, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Fisher, J.F.; Kavanagh, K.; Sobel, J.D.; Kauffman, C.A.; Newman, C.A. Candida urinary tract infection: Pathogenesis. Clin. Infect. Dis. 2011, 52, S437–S451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Cassone, A. Vulvovaginal Candida albicans infections: Pathogenesis, immunity and vaccine prospects. Int. J. Obstet. Gynaecol. 2015, 122, 785–794. [Google Scholar] [CrossRef]
  6. Cheng, S.; Clancy, C.J.; Checkley, M.A.; Handfield, M.; Hillman, J.D.; Progulske-Fox, A.; Lewin, A.S.; Fidel, P.L.; Nguyen, M.H. Identification of Candida albicans genes induced during thrush offers insight into pathogenesis. Mol. Microbiol. 2003, 48, 1275–1288. [Google Scholar] [CrossRef] [Green Version]
  7. Pappas, P.G.; Lionakis, M.S.; Arendrup, M.C.; Ostrosky-Zeichner, L.; Kullberg, B.J. Invasive candidiasis. Nat. Rev. Dis. Primers 2018, 4, 1–20. [Google Scholar] [CrossRef] [PubMed]
  8. Perlin, D.S. Mechanisms of echinocandin antifungal drug resistance. Ann. N. Y. Acad. Sci. 2015, 1354, 1. [Google Scholar] [CrossRef]
  9. Healey, K.R.; Perlin, D.S. Fungal resistance to echinocandins and the MDR phenomenon in Candida glabrata. J. Fungi 2018, 4, 105. [Google Scholar]
  10. Healey, K.R.; Zhao, Y.; Perez, W.B.; Lockhart, S.R.; Sobel, J.D.; Farmakiotis, D.; Kontoyiannis, D.P.; Sanglard, D.; Taj-Aldeen, S.J.; Alexander, B.D. Prevalent mutator genotype identified in fungal pathogen Candida glabrata promotes multi-drug resistance. Nat. Commnu. 2016, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  11. Kordalewska, M.; Perlin, D.S. Identification of drug resistant Candida auris. Front. Microbiol. 2019, 10, 1918. [Google Scholar] [CrossRef] [PubMed]
  12. Revie, N.M.; Iyer, K.R.; Robbins, N.; Cowen, L.E. Antifungal drug resistance: Evolution, mechanisms and impact. Curr. Opin. Microbiol. 2018, 45, 70–76. [Google Scholar] [CrossRef] [PubMed]
  13. Nantel, A.; Dignard, D.; Bachewich, C.; Harcus, D.; Marcil, A.; Bouin, A.-P.; Sensen, C.W.; Hogues, H.; van het Hoog, M.; Gordon, P. Transcription profiling of Candida albicans cells undergoing the yeast-to-hyphal transition. Mol. Biol. Cell 2002, 13, 3452–3465. [Google Scholar] [CrossRef] [Green Version]
  14. Whiteway, M.; Bachewich, C. Morphogenesis in Candida albicans. Annu. Rev. Microbiol. 2007, 61, 529–553. [Google Scholar] [CrossRef] [Green Version]
  15. Kumar, P.; Tan, Y.; Cahan, P. Understanding development and stem cells using single cell-based analyses of gene expression. Development 2017, 144, 17–32. [Google Scholar] [CrossRef] [Green Version]
  16. Rué, P.; Martinez Arias, A. Cell dynamics and gene expression control in tissue homeostasis and development. Mol. Syst. Biol. 2015, 11, 792. [Google Scholar] [CrossRef] [PubMed]
  17. Cardoso-Moreira, M.; Halbert, J.; Valloton, D.; Velten, B.; Chen, C.; Shao, Y.; Liechti, A.; Ascenção, K.; Rummel, C.; Ovchinnikova, S. Gene expression across mammalian organ development. Nature 2019, 571, 505–509. [Google Scholar] [CrossRef]
  18. Chandrangsu, P.; Rensing, C.; Helmann, J.D. Metal homeostasis and resistance in bacteria. Nat. Rev. Microbiol. 2017, 15, 338. [Google Scholar]
  19. Li, C.; Li, Y.; Ding, C. The role of copper homeostasis at the host-pathogen axis: From bacteria to fungi. Int. J. Mol. Sci. 2019, 20, 175. [Google Scholar] [CrossRef] [Green Version]
  20. Brickner, J.H.; Walter, P. Gene recruitment of the activated INO1 locus to the nuclear membrane. PLoS Biol. 2004, 2, e342. [Google Scholar] [CrossRef]
  21. Egecioglu, D.; Brickner, J.H. Gene positioning and expression. Curr. Opin. Cell Biol. 2011, 23, 338–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gasch, A.P.; Werner-Washburne, M. The genomics of yeast responses to environmental stress and starvation. Funct. Integr. Genom. 2002, 2, 181–192. [Google Scholar] [CrossRef] [PubMed]
  23. Gasch, A.P.; Yu, F.B.; Hose, J.; Escalante, L.E.; Place, M.; Bacher, R.; Kanbar, J.; Ciobanu, D.; Sandor, L.; Grigoriev, I.V. Single-cell RNA sequencing reveals intrinsic and extrinsic regulatory heterogeneity in yeast responding to stress. PLoS Biol. 2017, 15, e2004050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pérez-Martínez, M.E.; Benet, M.; Alepuz, P.; Tordera, V. Nut1/Hos1 and Sas2/Rpd3 control the H3 acetylation of two different sets of osmotic stress-induced genes. Epigenetics 2020, 15, 251–271. [Google Scholar] [CrossRef] [PubMed]
  25. Nakajima, E.; Shimaji, K.; Umegawachi, T.; Tomida, S.; Yoshida, H.; Yoshimoto, N.; Izawa, S.; Kimura, H.; Yamaguchi, M. The Histone deacetylase gene Rpd3 is required for starvation stress resistance. PLoS ONE 2016, 11, e0167554. [Google Scholar] [CrossRef]
  26. Rocha, C.R.; Schroppel, K.; Harcus, D.; Marcil, A.; Dignard, D.; Taylor, B.N.; Thomas, D.Y.; Whiteway, M.; Leberer, E. Signaling through adenylyl cyclase is essential for hyphal growth and virulence in the pathogenic fungus Candida albicans. Mol. Biol. Cell 2001, 12, 3631–3643. [Google Scholar] [CrossRef] [Green Version]
  27. Harcus, D.; Nantel, A.; Marcil, A.; Rigby, T.; Whiteway, M. Transcription profiling of cyclic AMP signaling in Candida albicans. Mol. Biol. Cell 2004, 15, 4490–4499. [Google Scholar] [CrossRef] [Green Version]
  28. Gasch, A.P.; Spellman, P.T.; Kao, C.M.; Carmel-Harel, O.; Eisen, M.B.; Storz, G.; Botstein, D.; Brown, P.O. Genomic expression programs in the response of yeast cells to environmental changes. Mol. Biol. Cell 2000, 11, 4241–4257. [Google Scholar] [CrossRef]
  29. Rangel, D.E. Stress induced cross-protection against environmental challenges on prokaryotic and eukaryotic microbes. World J. Microbiol. Biotechnol. 2011, 27, 1281–1296. [Google Scholar] [CrossRef]
  30. Kreuzer, K.N. DNA damage responses in prokaryotes: Regulating gene expression, modulating growth patterns, and manipulating replication forks. Cold Spring Harbor Perspect. Biol. 2013, 5, a012674. [Google Scholar] [CrossRef]
  31. Enjalbert, B.; Nantel, A.; Whiteway, M. Stress-induced gene expression in Candida albicans: Absence of a general stress response. Mol. Biol. Cell 2003, 14, 1460–1467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Leach, M.D.; Farrer, R.A.; Tan, K.; Miao, Z.; Walker, L.A.; Cuomo, C.A.; Wheeler, R.T.; Brown, A.J.; Wong, K.H.; Cowen, L.E. Hsf1 and Hsp90 orchestrate temperature-dependent global transcriptional remodelling and chromatin architecture in Candida albicans. Nat. Commun. 2016, 7, 1–13. [Google Scholar]
  33. Nobile, C.J.; Fox, E.P.; Nett, J.E.; Sorrells, T.R.; Mitrovich, Q.M.; Hernday, A.D.; Tuch, B.B.; Andes, D.R.; Johnson, A.D. A recently evolved transcriptional network controls biofilm development in Candida albicans. Cell 2012, 148, 126–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Fox, E.M.; Howlett, B.J. Secondary metabolism: Regulation and role in fungal biology. Curr. Opin. Microbiol. 2008, 11, 481–487. [Google Scholar] [CrossRef]
  35. Eldabagh, R.S.; Mejia, N.G.; Barrett, R.L.; Monzo, C.R.; So, M.K.; Foley, J.J.; Arnone, J.T. Systematic identification, characterization, and conservation of adjacent-gene coregulation in the budding yeast Saccharomyces cerevisiae. Msphere 2018, 3. [Google Scholar] [CrossRef] [Green Version]
  36. Hagee, D.; Hardan, A.A.; Botero, J.; Arnone, J.T. Genomic Clustering within Functionally Related Gene Families in Ascomycota Fungi. Computat. Struct. Biotechnol. J. 2020. [Google Scholar] [CrossRef]
  37. Xu, H.; Liu, J.-J.; Liu, Z.; Li, Y.; Jin, Y.-S.; Zhang, J. Synchronization of stochastic expressions drives the clustering of functionally related genes. Sci. Adv. 2019, 5, eaax6525. [Google Scholar] [CrossRef] [Green Version]
  38. Tye, B.W.; Commins, N.; Ryazanova, L.V.; Wühr, M.; Springer, M.; Pincus, D.; Churchman, L.S. Proteotoxicity from aberrant ribosome biogenesis compromises cell fitness. Elife 2019, 8, e43002. [Google Scholar] [CrossRef]
  39. Ashburner, M.; Ball, C.A.; Blake, J.A.; Botstein, D.; Butler, H.; Cherry, J.M.; Davis, A.P.; Dolinski, K.; Dwight, S.S.; Eppig, J.T. Gene ontology: Tool for the unification of biology. Nat. Genet. 2000, 25, 25–29. [Google Scholar] [CrossRef] [Green Version]
  40. Consortium, G.O. The gene ontology resource: 20 years and still GOing strong. Nucl. Acids Res. 2019, 47, D330–D338. [Google Scholar]
  41. Skrzypek, M.S.; Binkley, J.; Binkley, G.; Miyasato, S.R.; Simison, M.; Sherlock, G. The Candida Genome Database (CGD): Incorporation of Assembly 22, systematic identifiers and visualization of high throughput sequencing data. Nucl. Acids Res. 2016, gkw924. [Google Scholar]
  42. Arnone, J.T.; Robbins-Pianka, A.; Arace, J.R.; Kass-Gergi, S.; McAlear, M.A. The adjacent positioning of co-regulated gene pairs is widely conserved across eukaryotes. BMC Genom. 2012, 13, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Maguire, S.L.; OhÉigeartaigh, S.S.; Byrne, K.P.; Schröder, M.S.; O’Gaora, P.; Wolfe, K.H.; Butler, G. Comparative genome analysis and gene finding in Candida species using CGOB. Mol. Biol. Evol. 2013, 30, 1281–1291. [Google Scholar]
  44. Fitzpatrick, D.A.; O’Gaora, P.; Byrne, K.P.; Butler, G. Analysis of gene evolution and metabolic pathways using the Candida Gene Order Browser. BMC Genom. 2010, 11, 290. [Google Scholar] [CrossRef] [Green Version]
  45. Cera, A.; Holganza, M.K.; Hardan, A.A.; Gamarra, I.; Eldabagh, R.S.; Deschaine, M.; Elkamhawy, S.; Sisso, E.M.; Foley, J.J.; Arnone, J.T. Functionally Related Genes Cluster into Genomic Regions That Coordinate Transcription at a Distance in Saccharomyces cerevisiae. Msphere 2019, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Sun, N.; Li, D.; Zhang, Y.; Killeen, K.; Groutas, W.; Calderone, R. Repurposing an inhibitor of ribosomal biogenesis with broad anti-fungal activity. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef]
  47. Dorsaz, S.; Snäkä, T.; Favre-Godal, Q.; Maudens, P.; Boulens, N.; Furrer, P.; Ebrahimi, S.N.; Hamburger, M.; Allémann, E.; Gindro, K. Identification and mode of action of a plant natural product targeting human fungal pathogens. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar]
  48. Harcus, D.; Dignard, D.; Lépine, G.; Askew, C.; Raymond, M.; Whiteway, M.; Wu, C. Comparative xylose metabolism among the ascomycetes C. albicans, S. stipitis and S. cerevisiae. PLoS ONE 2013, 8, e80733. [Google Scholar] [CrossRef] [Green Version]
  49. Arnone, J.T.; McAlear, M.A. Adjacent gene pairing plays a role in the coordinated expression of ribosome biogenesis genes MPP10 and YJR003C in Saccharomyces cerevisiae. Eukar. Cell 2011, 10, 43–53. [Google Scholar] [CrossRef] [Green Version]
  50. Arnone, J.T.; Arace, J.R.; Soorneedi, A.R.; Citino, T.T.; Kamitaki, T.L.; McAlear, M.A. Dissecting the cis and trans elements that regulate adjacent-gene coregulation in Saccharomyces cerevisiae. Eukar. Cell 2014, 13, 738–748. [Google Scholar] [CrossRef] [Green Version]
  51. Wade, C.H.; Umbarger, M.A.; McAlear, M.A. The budding yeast rRNA and ribosome biosynthesis (RRB) regulon contains over 200 genes. Yeast 2006, 23, 293–306. [Google Scholar] [CrossRef] [PubMed]
  52. Rendsvig, J.K.; Workman, C.T.; Hoof, J.B. Bidirectional histone-gene promoters in Aspergillus: Characterization and application for multi-gene expression. Fungal Biol. Biotechnol. 2019, 6, 1–14. [Google Scholar] [CrossRef]
  53. Xu, Z.; Wei, W.; Gagneur, J.; Perocchi, F.; Clauder-Münster, S.; Camblong, J.; Guffanti, E.; Stutz, F.; Huber, W.; Steinmetz, L.M. Bidirectional promoters generate pervasive transcription in yeast. Nature 2009, 457, 1033–1037. [Google Scholar] [CrossRef] [PubMed]
  54. Cohen, B.A.; Mitra, R.D.; Hughes, J.D.; Church, G.M. A computational analysis of whole-genome expression data reveals chromosomal domains of gene expression. Nat. Genet. 2000, 26, 183–186. [Google Scholar] [CrossRef] [PubMed]
  55. Quintero-Cadena, P.; Sternberg, P.W. Enhancer sharing promotes neighborhoods of transcriptional regulation across eukaryotes. G3 Genes Genom. Genet. 2016, 6, 4167–4174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Conservation of functionally related clusters in divergent fungal lineages. The functionally clustered gene pairings from Candida albicans were analyzed for conservation of the pairings. The heat map depicts the conservation with either yellow (conservation of the pairing) or black (no conservation of the pairing).
Figure 1. Conservation of functionally related clusters in divergent fungal lineages. The functionally clustered gene pairings from Candida albicans were analyzed for conservation of the pairings. The heat map depicts the conservation with either yellow (conservation of the pairing) or black (no conservation of the pairing).
Microorganisms 09 00276 g001
Figure 2. Many functionally clustered gene sets exhibit tighter transcriptional coregulation than random subgroupings of comparable sizes. Shown is the transcriptional correlation for every possible functional clustering arrangement that could have arisen through bootstrapping with replacement. The PCC was calculated for 10,000 iterations that represent every possible combination of clustering that could have evolved (of comparable size to the actual cluster for each set), and the frequency histograms are presented. The actual PCC for the clustered gene set within each family is marked with an asterisk.
Figure 2. Many functionally clustered gene sets exhibit tighter transcriptional coregulation than random subgroupings of comparable sizes. Shown is the transcriptional correlation for every possible functional clustering arrangement that could have arisen through bootstrapping with replacement. The PCC was calculated for 10,000 iterations that represent every possible combination of clustering that could have evolved (of comparable size to the actual cluster for each set), and the frequency histograms are presented. The actual PCC for the clustered gene set within each family is marked with an asterisk.
Microorganisms 09 00276 g002aMicroorganisms 09 00276 g002b
Figure 3. Functionally related genes congregate to transcriptionally permissive genomic loci. The Spearman’s correlation coefficient was calculated as a function of genomic distance globally (A) and for the functionally clustered gene set (B) and the logarithmic decay of best fit was determined for the datasets. These functions were overlaid (C,D) for clarity of comparison. The red trend line represents the genome-wide dataset and the blue represents the clustered functions.
Figure 3. Functionally related genes congregate to transcriptionally permissive genomic loci. The Spearman’s correlation coefficient was calculated as a function of genomic distance globally (A) and for the functionally clustered gene set (B) and the logarithmic decay of best fit was determined for the datasets. These functions were overlaid (C,D) for clarity of comparison. The red trend line represents the genome-wide dataset and the blue represents the clustered functions.
Microorganisms 09 00276 g003
Table 1. Gene Clustering within a Shared Molecular Process in Candida albicans.
Table 1. Gene Clustering within a Shared Molecular Process in Candida albicans.
Molecular ProcessGene Family SizeSingletonsClustersp-Value
Biofilm formation149132172.53 × 10−4
Carbohydrate metabolic process191168238.19 × 10−4
Cell adhesion686172.94 × 10−5
Cell budding46460n.s
Cell cycle4773751021.02 × 10−4
Cell development11311036.03 × 10−1
Cell wall organization1601600n.s.
Cellular homeostasis205183229.99 × 10−3
Cellular protein modification process5844521323.92 × 10−3
Cellular respiration898541.19 × 10−1
Conjugation63630n.s.
Cytokinesis918741.36 × 10−1
Cytoskeleton organization184157275.38 × 10−6
DNA metabolic process344290543.24 × 10−3
Filamentous growth6074701389.55 × 10−3
Generation of precursor metabolites and energy140119214.34 × 10−7
Growth of unicellular organism as a thread of attached cells888441.11 × 10−1
Hyphal growth979163.50 × 10−2
Interspecies interaction between organisms343300431.58 × 10−1
Lipid metabolic process286249371.44 × 10−2
Nucleus organization565243.68 × 10−3
Organelle organization10536823716.68 × 10−3
Pathogenesis275239367.33 × 10−3
Protein catabolic process220196241.44 × 10−2
Protein folding82820n.s.
Pseudohyphal growth413921.78 × 10−2
Regulation of biological process15008596416.73 × 10−1
Response to chemical8045912131.19 × 10−1
Response to drug406343635.53 × 10−2
Response to stress8606142466.24 × 10−2
Ribosome biogenesis305277285.29 × 10−2
RNA metabolic process7745512211.23 × 10−3
Signal transduction219203163.85 × 10−1
Translation245207382.45 × 10−5
Transport10606983625.45 × 10−2
Transposition440n.s.
Vesicle-mediated transport320290306.62 × 10−1
Vitamin metabolic process403468.64 × 10−7
Candida haploid genome: 6066 total open reading frames. n.s.—not significant.
Table 2. Pearson’s Correlation Coefficient of Functionally Related Genes.
Table 2. Pearson’s Correlation Coefficient of Functionally Related Genes.
Molecular ProcessFamilySingletonsClustered Set
Biofilm formation0.2820.2660.376
Carbohydrate metabolic process0.2120.2240.101
Cell adhesion0.2740.2830.370
Cell cycle0.2580.2560.260
Cell development0.2160.2150.433
Cellular homeostasis0.2830.2660.429
Cellular protein modification process0.2890.2760.331
Cellular respiration0.2460.2500.293
Cytokinesis0.2990.3000.100
Cytoskeleton organization0.2410.2620.116
DNA metabolic process0.2420.2600.148
Filamentous growth0.2810.2950.237
Generation of precursor metabolites and energy0.2370.2310.287
Growth of unicellular organism as a thread of attached cells0.2800.3010.017
Hyphal growth0.3080.3060.287
Interspecies interaction between organisms0.3480.3680.182
Lipid metabolic process0.3450.3270.481
Nucleus organization0.5400.5250.772
Organelle organization0.2850.3100.240
Pathogenesis0.3630.3810.243
Protein catabolic process0.3280.3290.352
Pseudohyphal growth0.2600.262−0.438
Regulation of biological process0.2950.3040.282
Response to chemical0.2690.2960.196
Response to drug0.2900.2980.253
Response to stress0.3160.3020.229
Ribosome biogenesis0.6800.6730.736
RNA metabolic process0.4160.4150.470
Signal transduction0.2880.2790.372
Translation0.5250.5310.582
Transport0.2640.2710.250
Vesicle-mediated transport0.3190.3130.368
Vitamin metabolic process0.2580.2270.641
Table 3. Orientation of Functional Clusters.
Table 3. Orientation of Functional Clusters.
Molecular ProcessDivergentTandemConvergent
Biofilm formation360
Carbohydrate metabolic process231
Cell adhesion020
Cell cycle17298
Cell development110
Cellular homeostasis352
Cellular protein modification process153816
Cellular respiration110
Cytokinesis020
Cytoskeleton organization544
DNA metabolic process9115
Filamentous growth204015
Generation of precursor metabolites and energy121
Growth of unicellular organism as a thread of attached cells020
Hyphal growth020
Interspecies interaction between organisms6125
Lipid metabolic process4141
Nucleus organization010
Organelle organization628952
Pathogenesis795
Protein catabolic process143
Pseudohyphal growth010
Regulation of biological process8718094
Response to chemical226326
Response to drug4204
Response to stress306633
Ribosome biogenesis662
RNA metabolic process504327
Signal transduction512
Translation586
Transport164627
Vesicle-mediated transport583
Vitamin metabolic process211
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Asfare, S.; Eldabagh, R.; Siddiqui, K.; Patel, B.; Kaba, D.; Mullane, J.; Siddiqui, U.; Arnone, J.T. Systematic Analysis of Functionally Related Gene Clusters in the Opportunistic Pathogen, Candida albicans. Microorganisms 2021, 9, 276. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9020276

AMA Style

Asfare S, Eldabagh R, Siddiqui K, Patel B, Kaba D, Mullane J, Siddiqui U, Arnone JT. Systematic Analysis of Functionally Related Gene Clusters in the Opportunistic Pathogen, Candida albicans. Microorganisms. 2021; 9(2):276. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9020276

Chicago/Turabian Style

Asfare, Sarah, Reem Eldabagh, Khizar Siddiqui, Bharvi Patel, Diellza Kaba, Julie Mullane, Umar Siddiqui, and James T. Arnone. 2021. "Systematic Analysis of Functionally Related Gene Clusters in the Opportunistic Pathogen, Candida albicans" Microorganisms 9, no. 2: 276. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9020276

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop