Next Article in Journal
Nano-Silicon Triggers Rapid Transcriptomic Reprogramming and Biochemical Defenses in Brassica napus Challenged with Sclerotinia sclerotiorum
Previous Article in Journal
An Evaluation of the OLM PneumID Real-Time Polymerase Chain Reaction to Aid in the Diagnosis of Pneumocystis Pneumonia
Previous Article in Special Issue
Sticking to the Subject: Multifunctionality in Microbial Adhesins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Moonlighting Proteins: Diverse Functions Found in Fungi

1
Department of Biological Sciences, University of Illinois at Chicago, Chicago, IL 60607, USA
2
Bay Path University, Longmeadow, MA 01106, USA
*
Author to whom correspondence should be addressed.
J. Fungi 2023, 9(11), 1107; https://doi.org/10.3390/jof9111107
Submission received: 20 September 2023 / Revised: 4 November 2023 / Accepted: 9 November 2023 / Published: 15 November 2023
(This article belongs to the Special Issue Multifunctional Fungal Proteins)

Abstract

:
Moonlighting proteins combine multiple functions in one polypeptide chain. An increasing number of moonlighting proteins are being found in diverse fungal taxa that vary in morphology, life cycle, and ecological niche. In this mini-review we discuss examples of moonlighting proteins in fungi that illustrate their roles in transcription and DNA metabolism, translation and RNA metabolism, protein folding, and regulation of protein function, and their interaction with other cell types and host proteins.

1. Introduction

In the million years since fungi split from other eukaryotes, they evolved a wide diversity of taxa with a variety of morphologies and sometimes complex life cycles, adapted to numerous ecological niches. They also developed diverse mechanisms of nutrient acquisition, including secreting digestive enzymes, symbiosis, or even parasitism. All these developments required the evolution of new protein functions. Moonlighting proteins have arisen throughout the evolutionary tree as a means of enabling a vast increase in protein functions, as well as a mechanism to coordinate many functions. A moonlighting protein is a single polypeptide chain with two or more physiologically relevant biochemical or biophysical functions [1]. Moonlighting proteins can have a second function by interacting with another protein, DNA, or RNA, or with different ligands or cofactors; by joining a multiprotein complex; or by being expressed in a different cell type. They can also have different functions at different times in development or stages of differentiation; in different locations within the cell, for example, in the cytoplasm or nucleus; or when secreted or bound to the cell surface.
As shown in the examples below and in Table 1, moonlighting proteins have been found in a variety of fungal taxa [2]. Not surprisingly, additional functions were often first found in model organisms that are used in genetic or biochemical studies including Saccharomyces cerevisiae (budding yeast), Neurospora crassa (filamentus fungi, mold), Schizosaccharomyces pombe (fission yeast), and Emericella nidulans (Aspergillus nidulans). A few examples were identified in Komagataella pastoris (formerly known as Pichia pastoris), a budding and spore-forming yeast that is widely used as a model organism and also as an expression system for protein production in research and biotech industries. Other moonlighting proteins were found in studies of fungi that play roles in health and disease or are used in industry or food production—Paracoccidioides brasiliensis is a dimorphic fungus that can cause paracoccidioidomycosis, Histoplasma capsulatum causes histoplasmosis, Cryptococcus neoformans can cause a pneumonia-like disease or meningitis in people with weakened immune systems, and Sporothrix schenckii causes sporotrichosis, a subcutaneous mycosis. In addition, several Candida species are a normal part of human gut flora, but can also act as opportunistic pathogens. Kluyveromyces lactis (formerly Saccharomyces lactis) is used in genetic research and also in industry for production of chymosin (rennet) for cheese production, and Tuber melanosporum has fruiting bodies that are eaten as truffles.
In this mini-review we discuss examples of moonlighting fungal proteins. Most of the proteins are either enzymes or chaperones with at least one additional function. For clarity, we are grouping the proteins into four groups based on whether the additional function plays a role in 1. affecting transcription, DNA metabolism, or DNA maintenance; 2. affecting translation or RNA metabolism; 3. affecting another protein’s function; or 4. interacting with host cells or proteins (Figure 1). Additional fungal moonlighting proteins are described in other papers in this collection, including surface proteases that have additional functions such as acting as adhesins to host proteins [72]; cell surface adhesins with additional functions [73]; and a protein with different functions in multiple protein complexes, Hal3 [38].

2. Examples of Moonlighting Proteins in Fungi

2.1. Affecting Transcription, DNA Metabolism, or DNA Maintenance

A variety of fungal enzymes have been found to have a second function in regulating transcription or in DNA metabolism (Table 1). Enzymes in carbohydrate, nitrogen, and sulfur metabolism, as well as kinases and chaperones, have been found to regulate transcription either by binding directly to DNA or by binding to and regulating transcription factors. S. cerevisiae hexokinase (Hxk2) and a homologue in K. lactis, galactokinase, phosphorylate hexoses in glycolysis and other pathways in carbohydrate metabolism. High cellular glucose concentrations cause Hxk2 to move to the nucleus where it interacts directly with Mig1, which binds to promoters of genes that are repressed by glucose [13]. Galactokinase is a transcriptional activator of GAL genes [5,6].
Examples in nitrogen metabolism include S. cerevisiae Arg5,6 and Ure2. Arg5,6 (N-acetyl-gamma phosphate reductase and acetyl glutamate kinase) in the arginine biosynthetic pathway binds to mitochondrial and nuclear DNA in vivo and acts as a regulator of transcription for several genes [10]. Ure2 has glutathione peroxidase activity and also functions in nitrogen catabolite repression by binding to transcriptional activators Gln3 and Gat1. Binding by Ure2 keeps them in the cytoplasm and prevents transcription of their target genes [19,20,21,22].
An example in sulfur metabolism was found in the Perigord black truffle, T. melanosporum (a filamentous mycorrhizal ascomycetous fungus). Phosphoadenosine phosphosulfate reductase is a thioredoxin that functions in sulfur assimilation, by converting activated sulfate (PAPS) to sulfite (adenosine 3′,5′-bisphosphate + sulfite + thioredoxin disulfide <-> 3′-phosphoadenylyl sulfate + thioredoxin). It also has an additional function as a transcription factor [8].
Several other types of enzymes or chaperones help the cell respond to changes in the environment by regulating the actions of transcription factors. S. cerevisiae Arg82 kinase phosphorylates inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) and inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4) and has a second function in stabilizing the transcription factors Arg80 and Mcm1 [11]. S. cerevisiae zuotin, a component of the ribosome-associated chaperone complex (RAC) that helps in folding of nascent polypeptide chains, is also an activator of Pdr1 transcription factor [23]. Superoxide dismutase breaks down superoxide (O2) to hydrogen peroxide (H2O2), and it also acts as a transcription factor. In response to H2O2, it regulates expression of genes involved in oxidative resistance and DNA damage repair. Treatment with reactive oxygen species (ROS) increases Sod1 binding to the promoters of rnr3 and gre2 [17,18]. S. cerevisiae glyceraldehyde 3-phosphate dehydrogenase (GAPDH), affects gene expression more generally. It binds to Sir2 and regulates gene silencing by influencing Sir2’s association with chromatin, especially near telomeres [12]. T. melanosporum Ask1 protein is part of the DASH complex, which binds to microtubules and kinetochores and regulates their association. It also interacts with DNA in a separate function as a transcriptional regulator [7].
Other enzymes and chaperones have additional functions in DNA damage repair and maintaining mitochondrial DNA. Lys20 homocitrate synthase in the lysine biosynthetic pathway assists in Ino80 accumulation at DNA breaks to aid in histone eviction for DNA damage repair [16]. S. cerevisiae Hsp60 chaperonin is a protein folding chaperone that has a second function in which it binds the template strand of active mtDNA ori sequences and is involved in the structure and transmission of mitochondrial DNA nucleoids to daughter cells [15]. S. cerevisiae acetohydroxyacid reductoisomerase, an enzyme in L-leucine, L-isoleucine, and L-valine amino acid biosynthesis, is also involved in maintaining mitochondrial DNA stability enzymes [9].
Moonlighting proteins can also be involved in both DNA and RNA metabolism. E. nidulans (A. nidulans) I-AniI is an mRNA maturase that is required for splicing of the intron in the cytochrome b (cobA) gene, which contains the coding sequence for the maturase. I-Anil stimulates the intrinsic ribozyme activity of the intron by binding to and stabilizing the three-dimensional structure of the RNA. I-Anil is also a homing endonuclease that introduces a double-strand break at a specific location in the cobA gene and helps in the insertion of a group I intron containing its own coding sequence [3,4].

2.2. Affecting Translation or RNA Metabolism

In many branches of the evolutionary tree, enzymes from intermediary metabolism, amino acid tRNA synthetases, and protein components of the ribosome have been found to have a second function in RNA metabolism, including splicing, folding, regulation of translation, cellular localization, or RNA lifetime (Table 1).
In S. cerevisiae, several protein components of the ribosome interact with RNA as a second function. The S14 protein of the small subunit (40S) binds to an RNA stem-loop and represses expression of rps14B pre-mRNA [29]. The L32 protein of the large subunit (60S) inhibits the splicing of the transcript of its own gene, rpl32 [27]. The L2 protein of the large subunit shortens the half-life of L2 mRNA [26]. The S28 component of the small subunit also shortens the half-life of its own mRNA by binding to a hairpin structure in 3′ UTR and also to the decapping machinery [30].
S. cerevisiae isocitrate dehydrogenase is an enzyme in the citric acid cycle. Both subunits 1 and 2 were found to bind to 5′-untranslated leaders of mitochondrial mRNAs [25]. S. cerevisiae leucyl-tRNA synthetase, which attaches L-leucine to tRNA(Leu) for use in protein synthesis, is involved in group I intron splicing [28]. The N. crassa tyrosyl tRNA synthetase promotes folding of the group I intron catalytic core [24].
The number of examples of enzymes with an additional function in RNA metabolism is likely to grow significantly. In recent years, high-throughput proteomics studies have identified many other fungal proteins that bind RNA [74,75,76,77,78,79,80]. For example, S. cerevisiae RNA-binding proteins are summarized in the RBP2Go database [81], and they include all of the enzymes in glycolysis and the citric acid cycle. For many of the enzymes, it is not known if RNA binding is indicative of a second function in translation, RNA stability, localization, or splicing, or another RNA function, or if the RNA binding could, instead, be involved in regulation of the enzyme function. More experiments are needed to clarify the role and mechanisms of interactions with RNA.

2.3. Affecting Another Protein’s Activity

In addition to regulating protein function through regulation of transcription or translation, the actions of the many biochemical pathways and processes within a cell are coordinated through protein–protein interactions. In some cases, moonlighting proteins with a catalytic function have a second function where they interact with other proteins to facilitate biogenesis, folding, or assembly of other proteins or protein complexes (Table 1). S. cerevisiae arginase catalyzes the reaction L-arginine + H2O => L-ornithine + urea for breakdown and the use of arginine as a nitrogen source. Arginase can also regulate arginine biosynthesis. It senses changes in cellular ornithine and arginine concentrations and responds by binding to and inactivating the first enzyme in arginine biosynthesis, ornithine transcarbamylase [34].
S. cerevisiae peroxiredoxins 1 and 2 have a peroxidase antioxidant function. In response to changing cellular redox conditions, they switch to act as general molecular chaperones that help proteins fold [40]. Other moonlighting enzymes function in the folding or assembly of specific proteins or complexes. Hansenula poymorpha (Pichia angusta) pyruvate carboxylase, an enzyme in gluconeogenesis, has a second function in the assembly of peroxisomal alcohol oxidase (AOX). It may mediate FAD binding to the AOX monomers in the cytoplasm, which enables them to be transported to the peroxisome where they can assemble into the active octameric form [31,32]. S. cerevisiae fructose 1,6-bisphosphate aldolase, an enzyme in glycolysis and gluconeogenesis, has been found to bind to the vacuolar H+-ATPase and is needed for its assembly. The aldolase catalytic activity is not needed for this function [37]. S. cerevisiae subunit 3 of succinate dehydrogenase is a part of respiratory complex II in the electron transport chain in mitochondria, and it also helps in the biogenesis and assembly of membrane-integral subunits of the TIM22 complex, a mitochondrial inner membrane translocase [43,44,45,46].
Another way proteins can affect other protein functions is through acting as a scaffold to bring proteins or enzymes together. The S. cerevisiae RACK1 component of the 40S (small) ribosomal subunit acts as a scaffold in cytoplasmic signal transduction pathways [41,42].
Proteins can also help the cell respond to changes in the environment by switching from one function to serving as a signaling molecule to promote a different cellular process. S. cerevisiae cytochrome c is an electron carrier protein component of the mitochondrial electron transport chain. When released from the mitochondria it has a second function of binding to apoptosis protease activation factor-1 (Apaf-1) to promote apoptosis [35]. P. pastoris phosphofructokinase, an enzyme in glycolysis, is required for microautophagy (vacuolar degradation of peroxisomes). The catalytic activity is not needed, and only the alpha subunit of phosphofructokinase plays this role [33]. S. cerevisiae enolases 1 and 2 are required for vacuole homotypic membrane fusion and protein trafficking to the vacuole, and their catalytic activity is not needed for this role [36].
Yeast (S. pombe, S. cerevisiae) Hal3 proteins are found in two different heteromultimeric enzymes. As part of phosphopantothenoylcysteine decarboxylase, Hal3 helps catalyze the decarboxylation of phosphopahtothenoyl L-cysteine in coenzyme A biosynthesis. It also acts as a regulatory subunit of a serine/threonine phosphatase, Pzh1 or Ppz1 [39,82]. The multiple functions of Hal3 are described in more detail in another paper in this collection by Casamayor and Arino [38].

2.4. Intracellular Proteins with an Additional Function on the Cell Surface

Intracellular proteins that have a second function when outside the cell and bound to the cell surface comprise another type of moonlighting protein found throughout the evolutionary tree [83,84,85]. Both commensal and pathogenic species use these proteins as adhesins to interact with host cells and proteins: for example, fibronectin, laminin, and other proteins in the extracellular matrix (Table 1). Microbial pathogens often employ moonlighting proteins to bind to host plasminogen, helping convert it to the active protease plasmin so that the protease can aid in invading host tissues [86,87,88]. In some cases, pathogens use moonlighting surface proteins to bind directly to proteins of the immune system and interfere with their activity. In this section we summarize examples of intracellular proteins moonlighting on the cell surface in several species. For additional examples and information, Arvizu-Rubio and coworkers have recently published a more extensive list of moonlighting proteins on the surface of pathogenic fungi [89].
Due to roles as both commensal organisms and as opportunistic pathogens, several groups have identified moonlighting proteins on the Candida cell surface, including enzymes in carbohydrate metabolism and protein synthesis, chaperones, and enzymes in redox homeostasis (reviewed in [90]). C. albicans glyceraldehyde 3-phosphate dehydrogenase (GAPDH) binds to fibronectin and laminin in the extracellular matrix [53]. Triosephosphate isomerase from C. albicans and C. glabrata binds to several extracellular matrix proteins [61]. Enolase from C. albicans, C. tropicalis, and S. cerevisiae binds to vitronectin, fibronectin, and plasminogen [63]. C. albicans glutathione reductase [54,55] and phosphotransferase [48,62] bind to plasma proteins. C. tropicalis fructose-bisphosphate aldolase and enolase 1 from glycolysis [64] bind fibronectin, vitronectin, and laminin. C. albicans heat shock protein Ssa2, a chaperone, binds to HTN3/histatin 5, a peptide from human saliva [57]. H. capsulatum Hsp60 chaperone [66], P. brasiliensis triose phosphate isomerase (glycolysis and gluconeogenesis) [69], and Talaromyces marneffei (Penicillium marneffei) GAPDH [71] were all found to act as adhesins. P. brasiliensis enolase (glycolysis and gluconeogenesis) binds fibronectin [67], and P. brasiliensis GAPDH binds to fibronectin, laminin, and type I collagen [68]. In Sporothrix schenckii, a chaperone, GroEL/Hsp60, binds to the extracellular matrix proteins laminin, elastin, fibrinogen, and fibronectin [70]. Satala and coworkers recently added phosphoglycerate mutase to the list of C. albicans proteins interacting with host ECM proteins vitronectin and fibronectin [61].
Crowe and coworkers studied C. albicans proteins that bind to human plasminogen and identified alcohol dehydrogenase, thiol-specific antioxidant protein, translation elongation, peroxisomal catalase, fructose bisphosphate aldolase, phosphoglycerate kinase, and glyceraldehyde 3-phosphate dehydrogenase [50]. Enolase was also found to bind to plasminogen by another group [52]. Ef-Tu elongation factor and glycerol 3-phosphate dehydrogenase bind to factor H as well as to plasminogen [51,56]. Stie and coworkers identified several proteins in Cryptococcus neoformans that aid in invasion by interacting with plasminogen, including subunits of ATP synthase, pyruvate kinase, phosphoglycerate kinase, glucose 6-phosphate isomerase, and nitric oxide dioxygenase [65].
Moonlighting proteins that directly affect the host immune system include elongation factor 2 and superoxide dismutase 3. Both proteins interact with contact system proteins HK, FXII, and prekallikrein (PPK) [48,49]. 6-phosphogluconate dehydrogenase catalyzes the oxidative decarboxylation of 6-phosphogluconate to ribulose 5-phosphate with concomitant reduction of NADP to NADPH in the pentose phosphate pathway and has also been found to bind to components of the contact system [47,48,49]. The high affinity glucose transporter 1 (Hgt1p), a transmembrane sugar transporter, binds complement regulator FH and C4BP, and protects cells from the host’s complement cascade by reduced formation of the potentially lytic terminal complex (TCC) [58].
Like the many enzymes and chaperones that have been found to bind RNA through proteomics studies, many other intracellular proteins have also been found on the cell surface through surface proteomics studies [91]. Many of these proteins might also have moonlighting functions on the cell surface.

3. Discussion

The fungal moonlighting proteins described above include a large variety of proteins and combinations of functions. Some combinations of functions enable cells to sense and respond to changes in the environment; for example, proteins involved in redox reactions might sense changes in the redox state of the cell and then switch to a new function. Proteins involved in carbohydrate or nitrogen metabolism are often important sensors for the cellular levels of these nutrients. In general, moonlighting proteins can also enable coordinated regulation or crosstalk between biochemical pathways and cellular processes, such as energy producing pathways and transcription or translation. In other cases, the two functions might not be connected but might just be a case of a new function arising and the cell making use of proteins that are already available.
While we cannot always determine which function was the original function of the protein family, many moonlighting proteins belong to ancient protein families, notably enzymes in glycolysis, the citric acid cycle, and protein synthesis, including ribosomal proteins. Evolution of a new function does not always require large changes in the amino acid sequence or protein structure [92]. In fact, a binding site for plasminogen can be just a few amino acid changes in a protein surface loop or C-terminus. Many of these proteins have orthologues that are moonlighting in other organisms [2]. Interestingly, in some cases the more ancient, catalytic function is the same and a second function is different, a kind of “semi-orthologue.”
There are still many questions about moonlighting proteins. In many cases, very little is known about the molecular mechanism for performing one or both functions and how a protein might either switch between functions or perform both functions simultaneously. Even models or three-dimensional structures of a protein in one conformation or one complex often do not provide enough information about how it performs all of its functions. Structures of proteins in different conformations and in complex with other proteins, DNA, RNA or other ligands or cofactors are needed. In addition, for most of the intracellular proteins that have a second function on the cell surface, it is not known how the proteins can be targeted to these two different cellular locations, although several noncanonical secretion pathways have been found to contribute to intracellular proteins being found on the cell surface (reviewed in [93]).
For some proteins, it is clear from mutagenic studies that a protein has a function in addition to its catalytic function, but the biochemical or biophysical function is not known. For example, the DeLuna lab showed that mutant forms of the S. cerevisiae Alt1 and Bat2 transaminases and the isoleucine/valine biosynthetic enzymes Ilv1 and Ilv2 that lack catalytic function can partially restore the gene knockout phenotype [94]. Detomasi and coworkers similarly showed that N. crassa CWR-1 is a C1 oxidizing chitin polysaccharide monooxygenase as well as having a function in allorecognition during formation of an interconnected mycelial network. Mutant enzymes that lack the catalytic function still function in allorecognition, likely through protein–protein interactions with CWR-2 [95]. Flores and coworkers showed that deletion of Yarrowia lipolytica N-acetylglucosamine kinase (Nag5) results not only in loss of kinase activity but also in altered levels of expression of enzymes of the N-acetylglucosamine (NAGA) utilization pathway (NAG genes). Complementation with mutant Nag5 proteins with greatly decreased kinase activity could restore the regulation of gene expression [96]. These results indicate that Nag5 has separate kinase and gene regulation functions, although how it regulates gene expression, for example, by binding to DNA or to a transcription factor, is not known.
The effects of interactions with additional macromolecules are often not clear. Many enzymes in intermediary metabolism in yeast and other organisms have been found to bind RNA, but in most cases, it is not clear if the interaction affects aspects of RNA function—translation, splicing, localization, or lifetime—or, perhaps, aspects of the enzyme function—catalysis, localization, or degradation.
With such variety of moonlighting proteins and functions, there are not shared characteristics that have been identified for predicting if a protein has moonlighting functions, although some characteristics and methods can be used to accurately predict individual functions. In general, many moonlighting proteins have similar characteristics to other proteins found in the same cellular locations [97]. Many of the new functions were identified by serendipity, but more recently, proteomics studies have identified many examples [98]. For example, proteomics studies that identified proteins that bind to plasminogen or to RNA have identified many of the proteins mentioned above as well as similar moonlighting proteins in other species.
The importance of fungi in health and disease, agriculture, and biotechnology has played a role in the increasing number of moonlighting functions that have been identified in recent years, and studies of the moonlighting functions might be beneficial in these fields. Several of the species mentioned above are important components of the human microbiome, especially the gut microbiome, sometimes as commensal species and sometimes as pathogens. Some of the moonlighting proteins mentioned above are also involved in interfering with therapeutics or treatments, including adhesion to medical implants [99]. Understanding the functions of moonlighting proteins that interact with host proteins and act as adhesins, invasins, or immune system modulators could lead to improved therapeutics or preventative measures. Similarly, therapeutics or preventative measures might be found for treatment of fungal infections of agricultural animals or plants. In contrast, some fungi are important as food or in the production of food molecules, so understanding moonlighting proteins that affect cellular processes such as translation and protein production could be used in improving yields. Moonlighting proteins that sense small molecules and affect translation or transcription could be important starting points for synthetic biotechnology, for example, as sensors that affect translation or transcription.

4. Conclusions

Moonlighting proteins in fungi have a variety of functions and combinations of functions. In recent years, the use of proteomics methods to identify cell surface localized proteins or proteins that bind to DNA or RNA has contributed to the identification of dozens of additional moonlighting, or potentially moonlighting, proteins. The examples described above are likely to be only the tip of the iceberg, with additional functions of many more proteins still to be identified. There is still much to learn about the mechanisms of their functions and their cellular roles, with potential impacts in medicine, agriculture, and synthetic biology.

Author Contributions

Conceptualization, C.J.J.; resources, C.J.J. and N.J.C.; writing—original draft preparation, C.J.J., N.J.C., K.J.P. and A.R.; writing—review and editing, C.J.J., N.J.C., K.J.P. and A.R.; supervision, C.J.J.; project administration, C.J.J.; funding acquisition, C.J.J. and K.J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NSF grant numbers 2321442 and 2133816, a UIC Honors College grant, and a Chancellor’s Undergraduate Research Award.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the result.

References

  1. Jeffery, C.J. Moonlighting proteins. Trends Biochem. Sci. 1999, 24, 8–11. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, C.; Liu, H.; Zabad, S.; Rivera, N.; Rowin, E.; Hassan, M.; Gomez De Jesus, S.M.; Llinas Santos, P.S.; Kravchenko, K.; Mikhova, M.; et al. MoonProt 3.0: An update of the moonlighting proteins database. Nucleic Acids Res. 2020, 49, D368–D372. [Google Scholar] [CrossRef]
  3. Bolduc, J.M.; Spiegel, P.C.; Chatterjee, P.; Brady, K.L.; Downing, M.E.; Caprara, M.G.; Waring, R.B.; Stoddard, B.L. Structural and biochemical analyses of DNA and RNA binding by a bifunctional homing endonuclease and group I intron splicing factor. Genes Dev. 2003, 17, 2875–2888. [Google Scholar] [CrossRef]
  4. Ho, Y.; Kim, S.J.; Waring, R.B. A protein encoded by a group I intron in Aspergillus nidulans directly assists RNA splicing and is a DNA endonuclease. Proc. Natl. Acad. Sci. USA 1997, 94, 8994–8999. [Google Scholar] [CrossRef] [PubMed]
  5. Meyer, J.; Walker-Jonah, A.; Hollenberg, C.P. Galactokinase encoded by GAL1 is a bifunctional protein required for induction of the GAL genes in Kluyveromyces lactis and is able to suppress the gal3 phenotype in Saccharomyces cerevisiae. Mol. Cell. Biol. 1991, 11, 5454–5461. [Google Scholar] [PubMed]
  6. Bhat, P.J.; Oh, D.; Hopper, J.E. Analysis of the GAL3 signal transduction pathway activating GAL4 protein dependent transcription in Saccharomyces cerevisiae. Genetics 1990, 125, 281–291. [Google Scholar] [CrossRef] [PubMed]
  7. Montanini, B.; Levati, E.; Bolchi, A.; Kohler, A.; Morin, E.; Tisserant, E.; Martin, F.; Ottonello, S. Genome-wide search and functional identification of transcription factors in the mycorrhizal fungus Tuber melanosporum. New Phytol. 2011, 189, 736–750. [Google Scholar] [CrossRef]
  8. Levati, E.; Sartini, S.; Bolchi, A.; Ottonello, S.; Montanini, B. Moonlighting transcriptional activation function of a fungal sulfur metabolism enzyme. Sci. Rep. 2016, 6, 25165. [Google Scholar] [CrossRef]
  9. Bateman, J.M.; Perlman, P.S.; Butow, R.A. Mutational bisection of the mitochondrial DNA stability and amino acid biosynthetic functions of ilv5p of budding yeast. Genetics 2002, 161, 1043–1052. [Google Scholar] [CrossRef]
  10. Hall, D.A.; Zhu, H.; Zhu, X.; Royce, T.; Gerstein, M.; Snyder, M. Regulation of gene expression by a metabolic enzyme. Science 2004, 306, 482–484. [Google Scholar] [CrossRef]
  11. El Bakkoury, M.; Dubois, E.; Messenguy, F. Recruitment of the yeast MADS box proteins, ArgRI and Mcm1 by the pleiotropic factor ArgRIII is required for their stability. Mol. Microbiol. 2000, 35, 15–31. [Google Scholar] [CrossRef]
  12. Ringel, A.E.; Ryznar, R.; Picariello, H.; Huang, K.L.; Lazarus, A.G.; Holmes, S.G. Yeast Tdh3 (glyceraldehyde 3-phosphate dehydrogenase) is a Sir2-interacting factor that regulates transcriptional silencing and rDNA recombination. PLoS Genet. 2013, 9, e1003871. [Google Scholar] [CrossRef]
  13. Ahuatzi, D.; Herrero, P.; de la Cera, T.; Moreno, F. The glucose regulated nuclear localization of hexokinase 2 in Saccharomyces cerevisiae is Mig1-dependent. J. Biol. Chem. 2004, 279, 14440–14446. [Google Scholar] [CrossRef] [PubMed]
  14. Rodriguez, A.; De La Cera, T.; Herrero, P.; Moreno, F. The hexokinase 2 protein regulates the expression of the GLK1, HXK1 and HXK2 genes of Saccharomyces cerevisiae. Biochem. J. 2001, 355 Pt 3, 625–631. [Google Scholar] [CrossRef] [PubMed]
  15. Kaufman, B.A.; Kolesar, J.E.; Perlman, P.S.; Butow, R.A. A function for the mitochondrial chaperonin Hsp60 in the structure and transmission of mitochondrial DNA nucleoids in Saccharomyces cerevisiae. J. Cell. Biol. 2003, 163, 457–461. [Google Scholar] [CrossRef] [PubMed]
  16. Torres-Machorro, A.L.; Aris, J.P.; Pillus, L. A moonlighting metabolic protein influences repair at DNA double-stranded breaks. Nucleic Acids Res. 2015, 43, 1646–1658. [Google Scholar] [CrossRef]
  17. Bermingham-McDonogh, O.; Gralla, E.B.; Valentine, J.S. The copper, zinc superoxide dismutase gene of Saccharomyces cerevisiae: Cloning, sequencing, and biological activity. Proc. Natl. Acad. Sci. USA 1988, 85, 4789–4793. [Google Scholar] [CrossRef]
  18. Tsang, C.K.; Liu, Y.; Thomas, J.; Zhang, Y.; Zheng, X.F. Superoxide dismutase 1 acts as a nuclear transcription factor to regulate oxidative stress resistance. Nat. Commun. 2014, 5, 3446. [Google Scholar] [CrossRef]
  19. Coschigano, P.W.; Magasanik, B. The URE2 gene product of Saccharomyces cerevisiae plays an important role in the cellular response to the nitrogen source and has homology to glutathione S-transferases. Mol. Cell. Biol. 1991, 11, 822–832. [Google Scholar]
  20. Blinder, D.; Coschigano, P.W.; Magasanik, B. Interaction of the GATA factor Gln3p with the nitrogen regulator Ure2p in Saccharomyces cerevisiae. J. Bacteriol. 1996, 178, 4734–4736. [Google Scholar] [CrossRef]
  21. Courchesne, W.E.; Magasanik, B. Regulation of nitrogen assimilation in Saccharomyces cerevisiae: Roles of the URE2 and GLN3 genes. J. Bacteriol. 1988, 170, 708–713. [Google Scholar] [CrossRef]
  22. Bai, M.; Zhou, J.M.; Perrett, S. The yeast prion protein Ure2 shows glutathione peroxidase activity in both native and fibrillar forms. J. Biol. Chem. 2004, 279, 50025–50030. [Google Scholar] [CrossRef]
  23. Ducett, J.K.; Peterson, F.C.; Hoover, L.A.; Prunuske, A.J.; Volkman, B.F.; Craig, E.A. Unfolding of the C-terminal domain of the J-protein Zuo1 releases autoinhibition and activates Pdr1 dependent transcription. J. Mol. Biol. 2013, 425, 19–31. [Google Scholar] [CrossRef] [PubMed]
  24. Caprara, M.G.; Mohr, G.; Lambowitz, A.M. A tyrosyl-tRNA synthetase protein induces tertiary folding of the group I intron catalytic core. J. Mol. Biol. 1996, 257, 512–531. [Google Scholar] [CrossRef]
  25. Elzinga, S.D.; Bednarz, A.L.; van Oosterum, K.; Dekker, P.J.; Grivell, L.A. Yeast mitochondrial NAD(+)-dependent isocitrate dehydrogenase is an RNA-binding protein. Nucleic Acids Res. 1993, 21, 5328–5331. [Google Scholar] [CrossRef] [PubMed]
  26. Presutti, C.; Ciafre, S.A.; Bozzoni, I. The ribosomal protein L2 in S. cerevisiae controls the level of accumulation of its own mRNA. EMBO J. 1991, 10, 2215–2221. [Google Scholar] [CrossRef] [PubMed]
  27. Eng, F.J.; Warner, J.R. Structural basis for the regulation of splicing of a yeast messenger RNA. Cell 1991, 65, 797–804. [Google Scholar] [CrossRef]
  28. Sarkar, J.; Poruri, K.; Boniecki, M.T.; McTavish, K.K.; Martinis, S.A. Yeast mitochondrial leucyl-tRNA synthetase CP1 domain has functionally diverged to accommodate RNA splicing at expense of hydrolytic editing. J. Biol. Chem. 2012, 287, 14772–14781. [Google Scholar] [CrossRef]
  29. Fewell, S.W.; Woolford, J.L., Jr. Ribosomal protein S14 of Saccharomyces cerevisiae regulates its expression by binding to RPS14B pre-mRNA and to 18S rRNA. Mol. Cell. Biol. 1999, 19, 826–834. [Google Scholar] [CrossRef]
  30. Badis, G.; Saveanu, C.; Fromont-Racine, M.; Jacquier, A. Targeted mRNA degradation by deadenylation-independent decapping. Mol. Cell 2004, 15, 5–15. [Google Scholar] [CrossRef]
  31. Ozimek, P.; van Dijk, R.; Latchev, K.; Gancedo, C.; Wang, D.Y.; van der Klei, I.J.; Veenhuis, M. Pyruvate carboxylase is an essential protein in the assembly of yeast peroxisomal oligomeric alcohol oxidase. Mol. Biol. Cell 2003, 14, 786–797. [Google Scholar] [CrossRef]
  32. Ozimek, P.Z.; Klompmaker, S.H.; Visser, N.; Veenhuis, M.; van der Klei, I.J. The transcarboxylase domain of pyruvate carboxylase is essential for assembly of the peroxisomal flavoenzyme alcohol oxidase. FEMS Yeast Res. 2007, 7, 1082–1092. [Google Scholar] [CrossRef]
  33. Yuan, W.; Tuttle, D.L.; Shi, Y.J.; Ralph, G.S.; Dunn, W.A., Jr. Glucose-induced microautophagy in Pichia pastoris requires the alpha-subunit of phosphofructokinase. J. Cell Sci. 1997, 110 Pt 16, 1935–1945. [Google Scholar] [CrossRef]
  34. Bechet, J.; Wiame, J.M. Indication of a specific regulatory binding protein for ornithinetranscarbamylase in Saccharomyces cerevisiae. Biochem. Biophys. Res. Commun. 1965, 21, 226–234. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, T.; Wang, X.; Purring-Koch, C.; Wei, Y.; McLendon, G.L. A mutational epitope for cytochrome C binding to the apoptosis protease activation factor-1. J. Biol. Chem. 2001, 276, 13034–13038. [Google Scholar] [CrossRef] [PubMed]
  36. Decker, B.L.; Wickner, W.T. Enolase activates homotypic vacuole fusion and protein transport to the vacuole in yeast. J. Biol. Chem. 2006, 281, 14523–14528. [Google Scholar] [CrossRef] [PubMed]
  37. Lu, M.; Holliday, L.S.; Zhang, L.; Dunn, W.A., Jr.; Gluck, S.L. Interaction between aldolase and vacuolar H+-ATPase: Evidence for direct coupling of glycolysis to the ATP-hydrolyzing proton pump. J. Biol. Chem. 2001, 276, 30407–30413. [Google Scholar] [CrossRef] [PubMed]
  38. Casamayor, A.; Arino, J. Fungal Hal3 (and Its Close Relative Cab3) as Moonlighting Proteins. J. Fungi. 2022, 8, 1066. [Google Scholar] [CrossRef]
  39. Ruiz, A.; Gonzalez, A.; Munoz, I.; Serrano, R.; Abrie, J.A.; Strauss, E.; Arino, J. Moonlighting proteins Hal3 and Vhs3 form a heteromeric PPCDC with Ykl088w in yeast CoA biosynthesis. Nat. Chem. Biol. 2009, 5, 920–928. [Google Scholar] [CrossRef]
  40. Jang, H.H.; Lee, K.O.; Chi, Y.H.; Jung, B.G.; Park, S.K.; Park, J.H.; Lee, J.R.; Lee, S.S.; Moon, J.C.; Yun, J.W.; et al. Two enzymes in one; two yeast peroxiredoxins display oxidative stress-dependent switching from a peroxidase to a molecular chaperone function. Cell 2004, 117, 625–635. [Google Scholar] [CrossRef]
  41. Sengupta, J.; Nilsson, J.; Gursky, R.; Spahn, C.M.; Nissen, P.; Frank, J. Identification of the versatile scaffold protein RACK1 on the eukaryotic ribosome by cryo-EM. Nat. Struct. Mol. Biol. 2004, 11, 957–962. [Google Scholar] [CrossRef] [PubMed]
  42. Kadrmas, J.L.; Smith, M.A.; Pronovost, S.M.; Beckerle, M.C. Characterization of RACK1 function in Drosophila development. Dev. Dyn. 2007, 236, 2207–2215. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, F.; Huo, X.; Zhai, Y.; Wang, A.; Xu, J.; Su, D.; Bartlam, M.; Rao, Z. Crystal structure of mitochondrial respiratory membrane protein complex II. Cell 2005, 121, 1043–1057. [Google Scholar] [CrossRef] [PubMed]
  44. Lemire, B.D.; Oyedotun, K.S. The Saccharomyces cerevisiae mitochondrial succinate:ubiquinone oxidoreductase. Biochim. Biophys. Acta 2002, 1553, 102–116. [Google Scholar] [CrossRef]
  45. Yankovskaya, V.; Horsefield, R.; Tornroth, S.; Luna-Chavez, C.; Miyoshi, H.; Leger, C.; Byrne, B.; Cecchini, G.; Iwata, S. Architecture of succinate dehydrogenase and reactive oxygen species generation. Science 2003, 299, 700–704. [Google Scholar] [CrossRef]
  46. Gebert, N.; Gebert, M.; Oeljeklaus, S.; von der Malsburg, K.; Stroud, D.A.; Kulawiak, B.; Wirth, C.; Zahedi, R.P.; Dolezal, P.; Wiese, S.; et al. Dual function of Sdh3 in the respiratory chain and TIM22 protein translocase of the mitochondrial inner membrane. Mol. Cell 2011, 44, 811–818. [Google Scholar] [CrossRef] [PubMed]
  47. Beitner, R.; Nordenberg, J. Inhibition of 6-phosphogluconate dehydrogenase (decarboxylating) by glucose 1,6-bisphosphate. Biochim. Biophys. Acta 1979, 583, 266–269. [Google Scholar] [CrossRef] [PubMed]
  48. Seweryn, K.; Karkowska-Kuleta, J.; Wolak, N.; Bochenska, O.; Kedracka-Krok, S.; Kozik, A.; Rapala-Kozik, M. Kinetic and thermodynamic characterization of the interactions between the components of human plasma kinin-forming system and isolated and purified cell wall proteins of Candida albicans. Acta Biochim. Pol. 2015, 62, 825–835. [Google Scholar] [CrossRef]
  49. Karkowska-Kuleta, J.; Kedracka-Krok, S.; Rapala-Kozik, M.; Kamysz, W.; Bielinska, S.; Karafova, A.; Kozik, A. Molecular determinants of the interaction between human high molecular weight kininogen and Candida albicans cell wall: Identification of kininogen-binding proteins on fungal cell wall and mapping the cell wall binding regions on kininogen molecule. Peptides 2011, 32, 2488–2496. [Google Scholar] [CrossRef]
  50. Crowe, J.D.; Sievwright, I.K.; Auld, G.C.; Moore, N.R.; Gow, N.A.; Booth, N.A. Candida albicans binds human plasminogen: Identification of eight plasminogen-binding proteins. Mol. Microbiol. 2003, 47, 1637–1651. [Google Scholar] [CrossRef]
  51. Kunert, A.; Losse, J.; Gruszin, C.; Huhn, M.; Kaendler, K.; Mikkat, S.; Volke, D.; Hoffmann, R.; Jokiranta, T.S.; Seeberger, H.; et al. Immune evasion of the human pathogen Pseudomonas aeruginosa: Elongation factor Tuf is a factor H and plasminogen binding protein. J. Immunol. 2007, 179, 2979–2988. [Google Scholar] [CrossRef] [PubMed]
  52. Jong, A.Y.; Chen, S.H.M.; Stins, M.F.; Kim, K.S.; Tuan, T.L.; Huang, S.H. Binding of Candida albicans enolase to plasmin(ogen) results in enhanced invasion of human brain microvascular endothelial cells. J. Med. Microbiol. 2003, 52 Pt 8, 615–622. [Google Scholar] [CrossRef] [PubMed]
  53. Gozalbo, D.; Gil-Navarro, I.; Azorin, I.; Renau-Piqueras, J.; Martinez, J.P.; Gil, M.L. The cell wall-associated glyceraldehyde-3-phosphate dehydrogenase of Candida albicans is also a fibronectin and laminin binding protein. Infect. Immun. 1998, 66, 2052–2059. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, X.; De Micheli, M.; Coleman, S.T.; Sanglard, D.; Moye-Rowley, W.S. Analysis of the oxidative stress regulation of the Candida albicans transcription factor, Cap1p. Mol. Microbiol. 2000, 36, 618–629. [Google Scholar] [CrossRef]
  55. Gil-Bona, A.; Amador-Garcia, A.; Gil, C.; Monteoliva, L. The external face of Candida albicans: A proteomic view of the cell surface and the extracellular environment. J. Proteom. 2018, 180, 70–79. [Google Scholar] [CrossRef]
  56. Luo, S.; Hoffmann, R.; Skerka, C.; Zipfel, P.F. Glycerol-3-phosphate dehydrogenase 2 is a novel factor H-, factor H-like protein 1-, and plasminogen-binding surface protein of Candida albicans. J. Infect. Dis. 2013, 207, 594–603. [Google Scholar] [CrossRef]
  57. Cuéllar-Cruz, M.; Gutiérrez-Sánchez, G.; López-Romero, E.; Ruiz-Baca, E.; Villagómez-Castro, J.; Rodríguez-Sifuentes, L. Identification of Candida albicans heat shock proteins and Candida glabrata and Candida krusei enolases involved in the response to oxidative stress. Open Life Sci. 2013, 8, 337–345. [Google Scholar] [CrossRef]
  58. Lesiak-Markowicz, I.; Vogl, G.; Schwarzmuller, T.; Speth, C.; Lass-Florl, C.; Dierich, M.P.; Kuchler, K.; Wurzner, R. Candida albicans Hgt1p, a multifunctional evasion molecule: Complement inhibitor, CR3 analogue, and human immunodeficiency virus-binding molecule. J. Infect. Dis. 2011, 204, 802–809. [Google Scholar] [CrossRef]
  59. Gilmore, B.J.; Retsinas, E.M.; Lorenz, J.S.; Hostetter, M.K. An iC3b receptor on Candida albicans: Structure, function, and correlates for pathogenicity. J. Infect. Dis. 1988, 157, 38–46. [Google Scholar] [CrossRef]
  60. Hostetter, M.K.; Lorenz, J.S.; Preus, L.; Kendrick, K.E. The iC3b receptor on Candida albicans: Subcellular localization and modulation of receptor expression by glucose. J. Infect. Dis. 1990, 161, 761–768. [Google Scholar] [CrossRef]
  61. Satala, D.; Zelazna, A.; Satala, G.; Bukowski, M.; Zawrotniak, M.; Rapala-Kozik, M.; Kozik, A. Towards understanding the novel adhesin function of Candida albicans phosphoglycerate mutase at the pathogen cell surface: Some structural analysis of the interactions with human host extracellular matrix proteins. Acta Biochim. Pol. 2021, 68, 515–525. [Google Scholar] [CrossRef]
  62. Gaudet, P.; Livstone, M.S.; Lewis, S.E.; Thomas, P.D. Phylogenetic-based propagation of functional annotations within the Gene Ontology consortium. Brief. Bioinform. 2011, 12, 449–462. [Google Scholar] [CrossRef] [PubMed]
  63. Satala, D.; Satala, G.; Karkowska-Kuleta, J.; Bukowski, M.; Kluza, A.; Rapala-Kozik, M.; Kozik, A. Structural Insights into the Interactions of Candidal Enolase with Human Vitronectin, Fibronectin and Plasminogen. Int. J. Mol. Sci. 2020, 21, 7843. [Google Scholar] [CrossRef]
  64. Kozik, A.; Karkowska-Kuleta, J.; Zajac, D.; Bochenska, O.; Kedracka-Krok, S.; Jankowska, U.; Rapala-Kozik, M. Fibronectin-, vitronectin- and laminin-binding proteins at the cell walls of Candida parapsilosis and Candida tropicalis pathogenic yeasts. BMC Microbiol. 2015, 15, 197. [Google Scholar] [CrossRef]
  65. Stie, J.; Bruni, G.; Fox, D. Surface-associated plasminogen binding of Cryptococcus neoformans promotes extracellular matrix invasion. PLoS ONE 2009, 4, e5780. [Google Scholar] [CrossRef] [PubMed]
  66. Wampler, J.L.; Kim, K.P.; Jaradat, Z.; Bhunia, A.K. Heat shock protein 60 acts as a receptor for the Listeria adhesion protein in Caco-2 cells. Infect. Immun. 2004, 72, 931–936. [Google Scholar] [CrossRef] [PubMed]
  67. Donofrio, F.C.; Calil, A.C.A.; Miranda, E.T.; Almeida, A.M.F.; Benard, G.; Soares, C.P.; Veloso, S.N.; Soares, C.M.A.; Mendes Giannini, M.J.S. Enolase from Paracoccidioides brasiliensis: Isolation and identification as a fibronectin-binding protein. J. Med. Microbiol. 2009, 58 Pt 6, 706–713. [Google Scholar] [CrossRef]
  68. Barbosa, M.S.; Bao, S.N.; Andreotti, P.F.; de Faria, F.P.; Felipe, M.S.; dos Santos Feitosa, L.; Mendes-Giannini, M.J.; Soares, C.M. Glyceraldehyde-3-phosphate dehydrogenase of Paracoccidioides brasiliensis is a cell surface protein involved in fungal adhesion to extracellular matrix proteins and interaction with cells. Infect. Immun. 2006, 74, 382–389. [Google Scholar] [CrossRef]
  69. Pereira, L.A.; Bao, S.N.; Barbosa, M.S.; da Silva, J.L.; Felipe, M.S.; de Santana, J.M.; Mendes-Giannini, M.J.; de Almeida Soares, C.M. Analysis of the Paracoccidioides brasiliensis triosephosphate isomerase suggests the potential for adhesin function. FEMS Yeast Res. 2007, 7, 1381–1388. [Google Scholar] [CrossRef]
  70. Garcia-Carnero, L.C.; Salinas-Marin, R.; Lozoya-Perez, N.E.; Wrobel, K.; Wrobel, K.; Martinez-Duncker, I.; Nino-Vega, G.A.; Mora-Montes, H.M. The Heat Shock Protein 60 and Pap1 Participate in the Sporothrixschenckii-Host Interaction. J. Fungi. 2021, 7, 960. [Google Scholar] [CrossRef]
  71. Lau, S.K.; Tse, H.; Chan, J.S.; Zhou, A.C.; Curreem, S.O.; Lau, C.C.; Yuen, K.Y.; Woo, P.C. Proteome profiling of the dimorphic fungus Penicillium marneffei extracellular proteins and identification of glyceraldehyde-3-phosphate dehydrogenase as an important adhesion factor for conidial attachment. FEBS J. 2013, 280, 6613–6626. [Google Scholar] [CrossRef] [PubMed]
  72. Satala, D.; Bras, G.; Kozik, A.; Rapala-Kozik, M.; Karkowska-Kuleta, J. More than Just Protein Degradation: The Regulatory Roles and Moonlighting Functions of Extracellular Proteases Produced by Fungi Pathogenic for Humans. J. Fungi. 2023, 9, 121. [Google Scholar] [CrossRef] [PubMed]
  73. Lipke, P.N.; Ragonis-Bachar, P. Sticking to the Subject: Multifunctionality in Microbial Adhesins. J. Fungi. 2023, 9, 419. [Google Scholar] [CrossRef] [PubMed]
  74. Scherrer, T.; Mittal, N.; Janga, S.C.; Gerber, A.P. A screen for RNA-binding proteins in yeast indicates dual functions for many enzymes. PLoS ONE 2010, 5, e15499. [Google Scholar] [CrossRef] [PubMed]
  75. Tsvetanova, N.G.; Klass, D.M.; Salzman, J.; Brown, P.O. Proteome-wide search reveals unexpected RNA-binding proteins in Saccharomyces cerevisiae. PLoS ONE 2010, 5, e12671. [Google Scholar] [CrossRef]
  76. Kramer, K.; Sachsenberg, T.; Beckmann, B.M.; Qamar, S.; Boon, K.L.; Hentze, M.W.; Kohlbacher, O.; Urlaub, H. Photo-cross-linking and high-resolution mass spectrometry for assignment of RNA-binding sites in RNA-binding proteins. Nat. Methods 2014, 11, 1064–1070. [Google Scholar] [CrossRef]
  77. Beckmann, B.M.; Horos, R.; Fischer, B.; Castello, A.; Eichelbaum, K.; Alleaume, A.M.; Schwarzl, T.; Curk, T.; Foehr, S.; Huber, W.; et al. The RNA-binding proteomes from yeast to man harbour conserved enigmRBPs. Nat. Commun. 2015, 6, 10127. [Google Scholar] [CrossRef] [PubMed]
  78. Matia-Gonzalez, A.M.; Laing, E.E.; Gerber, A.P. Conserved mRNA-binding proteomes in eukaryotic organisms. Nat. Struct. Mol. Biol. 2015, 22, 1027–1033. [Google Scholar] [CrossRef]
  79. Shchepachev, V.; Bresson, S.; Spanos, C.; Petfalski, E.; Fischer, L.; Rappsilber, J.; Tollervey, D. Defining the RNA interactome by total RNA-associated protein purification. Mol. Syst. Biol. 2019, 15, e8689. [Google Scholar] [CrossRef]
  80. Curtis, N.J.; Jeffery, C.J. The expanding world of metabolic enzymes moonlighting as RNA binding proteins. Biochem. Soc. Trans. 2021, 49, 1099–1108. [Google Scholar] [CrossRef]
  81. Caudron-Herger, M.; Jansen, R.E.; Wassmer, E.; Diederichs, S. RBP2GO: A comprehensive pan-species database on RNA-binding proteins, their interactions and functions. Nucleic Acids Res. 2020, 49, D425–D436. [Google Scholar] [CrossRef] [PubMed]
  82. Molero, C.; Petrenyi, K.; Gonzalez, A.; Carmona, M.; Gelis, S.; Abrie, J.A.; Strauss, E.; Ramos, J.; Dombradi, V.; Hidalgo, E.; et al. The Schizosaccharomyces pombe fusion gene hal3 encodes three distinct activities. Mol. Microbiol. 2013, 90, 367–382. [Google Scholar] [CrossRef] [PubMed]
  83. Jeffery, C.J. Intracellular/surface moonlighting proteins that aid in the attachment of gut microbiota to the host. AIMS Microbiol. 2019, 5, 77–86. [Google Scholar] [CrossRef] [PubMed]
  84. Kainulainen, V.; Korhonen, T.K. Dancing to another tune-adhesive moonlighting proteins in bacteria. Biology 2014, 3, 178–204. [Google Scholar] [CrossRef]
  85. Jeffery, C. Intracellular proteins moonlighting as bacterial adhesion factors. AIMS Microbiol. 2018, 4, 362–376. [Google Scholar] [CrossRef]
  86. Raymond, B.B.; Djordjevic, S. Exploitation of plasmin(ogen) by bacterial pathogens of veterinary significance. Vet. Microbiol. 2015, 178, 1–13. [Google Scholar] [CrossRef]
  87. Collen, D.; Verstraete, M. Molecular biology of human plasminogen. II. Metabolism in physiological and some pathological conditions in man. Thromb. Diath. Haemorrh. 1975, 34, 403–408. [Google Scholar]
  88. Dano, K.; Andreasen, P.A.; Grondahl-Hansen, J.; Kristensen, P.; Nielsen, L.S.; Skriver, L. Plasminogen activators, tissue degradation, and cancer. Adv. Cancer Res. 1985, 44, 139–266. [Google Scholar]
  89. Arvizu-Rubio, V.J.; Garcia-Carnero, L.C.; Mora-Montes, H.M. Moonlighting proteins in medically relevant fungi. PeerJ 2022, 10, e14001. [Google Scholar] [CrossRef]
  90. Satala, D.; Karkowska-Kuleta, J.; Zelazna, A.; Rapala-Kozik, M.; Kozik, A. Moonlighting Proteins at the Candidal Cell Surface. Microorganisms 2020, 8, 1046. [Google Scholar] [CrossRef]
  91. Wang, W.; Jeffery, C.J. An analysis of surface proteomics results reveals novel candidates for intracellular/surface moonlighting proteins in bacteria. Mol. Biosyst. 2016, 12, 1420–1431. [Google Scholar] [CrossRef]
  92. Jeffery, C.J. Protein species and moonlighting proteins: Very small changes in a protein’s covalent structure can change its biochemical function. J. Proteom. 2016, 134, 19–24. [Google Scholar] [CrossRef]
  93. Miura, N.; Ueda, M. Evaluation of Unconventional Protein Secretion by Saccharomyces cerevisiae and other Fungi. Cells 2018, 7, 128. [Google Scholar] [CrossRef] [PubMed]
  94. Espinosa-Cantu, A.; Ascencio, D.; Herrera-Basurto, S.; Xu, J.; Roguev, A.; Krogan, N.J.; DeLuna, A. Protein Moonlighting Revealed by Noncatalytic Phenotypes of Yeast Enzymes. Genetics 2018, 208, 419–431. [Google Scholar] [CrossRef] [PubMed]
  95. Detomasi, T.C.; Rico-Ramirez, A.M.; Sayler, R.I.; Goncalves, A.P.; Marletta, M.A.; Glass, N.L. A moonlighting function of a chitin polysaccharide monooxygenase, CWR-1, in Neurospora crassa allorecognition. Elife 2022, 11, e80459. [Google Scholar] [CrossRef] [PubMed]
  96. Flores, C.L.; Arino, J.; Gancedo, C. The N-Acetylglucosamine Kinase from Yarrowia lipolytica Is a Moonlighting Protein. Int. J. Mol. Sci. 2021, 22, 13109. [Google Scholar] [CrossRef] [PubMed]
  97. Amblee, V.; Jeffery, C.J. Physical Features of Intracellular Proteins that Moonlight on the Cell Surface. PLoS ONE 2015, 10, e0130575. [Google Scholar] [CrossRef]
  98. Jeffery, C. The Use of Proteomics Studies in Identifying Moonlighting Proteins. Methods Mol. Biol. 2019, 1871, 437–443. [Google Scholar]
  99. Nunez-Beltran, A.; Lopez-Romero, E.; Cuellar-Cruz, M. Identification of proteins involved in the adhesionof Candida species to different medical devices. Microb. Pathog. 2017, 107, 293–303. [Google Scholar] [CrossRef]
Figure 1. Examples of multiple functions of fungal moonlighting proteins. A moonlighting protein (dark blue) can have a function as an enzyme in the cytoplasm (A). It could have an additional function as a DNA binding transcription factor in the nucleus (brown circle, B). By binding to RNA the moonlighting protein could affect splicing (C1), folding of the RNA (magenta, C2), and/or translation by the ribosome (C3). It could interact with another protein (red) to act as a chaperone to aid in protein folding (D1), or act as a scaffold that brings multiple proteins (green, pink, purple) together in a metabolon or as a subunit of a multiprotein complex (D2), or it could inhibit the function of another protein (yellow, D3). An intracellular protein could have an additional function on the cell surface where it interacts with host proteins, such as components of the extracellular matrix (purple and blue, E). Created with BioRender.com.
Figure 1. Examples of multiple functions of fungal moonlighting proteins. A moonlighting protein (dark blue) can have a function as an enzyme in the cytoplasm (A). It could have an additional function as a DNA binding transcription factor in the nucleus (brown circle, B). By binding to RNA the moonlighting protein could affect splicing (C1), folding of the RNA (magenta, C2), and/or translation by the ribosome (C3). It could interact with another protein (red) to act as a chaperone to aid in protein folding (D1), or act as a scaffold that brings multiple proteins (green, pink, purple) together in a metabolon or as a subunit of a multiprotein complex (D2), or it could inhibit the function of another protein (yellow, D3). An intracellular protein could have an additional function on the cell surface where it interacts with host proteins, such as components of the extracellular matrix (purple and blue, E). Created with BioRender.com.
Jof 09 01107 g001
Table 1. Moonlighting Proteins Mentioned in the Text.
Table 1. Moonlighting Proteins Mentioned in the Text.
Protein NameOne FunctionAnother FunctionReference Number
Interacting with DNA or affecting DNA metabolism
Emericella nidulans
I-AniI mRNA maturaseHoming endonuclease[3,4]
Kluyveromyces lactis
Galactokinase GalactokinaseTranscriptional activator[5,6]
Tuber melanosporum
Ask1 Component of Binds DNA,[7]
DASH complextranscriptional regulator
PhosphoadenosineThioredoxinTranscription factor [8]
phosphosulfate reductase
Saccharomyces cerevisiae
AcetohydroxyacidAcetohydroxyacidMaintains mitochondrial[9]
reductoisomerase reductoisomeraseDNA stability
Arg5,6N-acetyl-gammaRegulator of transcription[10]
phosphate reductase/acetyl glutamate kinase
Arg82KinaseStabilizes transcription factors[11]
ARG80 and MCM1
Glycerol 3-phosphateGlycerol 3-phosphateBinds to Sir2[12]
Dehydrogenasedehydrogenase
HexokinaseHexokinaseBinds Mig1[13,14]
Hsp60ChaperoninBinds DNA[15]
Lys20Homocitrate synthaseDNA damage repair[16]
Superoxide dismutaseSuperoxide dismutaseBinds DNA[17,18]
Ure2functions inGlutathione peroxidase, enzyme[19,20,21,22]
nitrogen catabolite repression
ZuotinComponent of theActivator of a transcription factor [23]
RAC chaperone complex
Binding to RNA
Neurospora crassa
Tyrosyl tRNA synthetaseCatalyzes aminoacylationPromotes folding of introns[24]
of tyrosine
Saccharomyces cerevisiae
Isocitrate dehydrogenase 1Isocitrate dehydrogenaseBinds mRNA [25]
Isocitrate dehydrogenase 2Isocitrate dehydrogenase Binds mRNA [25]
L2 Ribosomal proteinComponent of ribosomeShortens half-life of L2 mRNA[26]
L32 Ribosomal proteinComponent of ribosome Inhibits the splicing of RNA[27]
Leucyl-tRNA synthetaseLeucyl-tRNA synthetase Intron splicing[28]
S14 Ribosomal proteinComponent of ribosome Binds RNA, translation repressor[29]
S28 Ribosomal proteinComponent of ribosome Binds, shortens half-life of mRNA[30]
Protein–Protein Interactions
Hansenula poymorpha
Pyruvate carboxylasePyruvate carboxylaseAssembly of peroxisomal[31,32]
alcohol oxidase
Pichia pastoris
PhosphofructokinasePhosphofructokinaseMicroautophagy[33]
Saccharomyces cerevisiae
ArginaseArginaseBinds and inactivates[34]
ornithine transcarbamylase
Cytochrome CElectron carrier proteinBinds apoptosis proteins[35]
in electron transport chain
Enolase 1Enolase Vacuole-membrane fusion[36]
and protein trafficking to the vacuole
Enolase 2Enolase Vacuole-membrane fusion[36]
and protein trafficking to the vacuole
Fructose 1,6-bisphosphate Aldolase Binds vacuolar H+-ATPase[37]
aldolase and is needed for its assembly
Hal3Subunit ofInhibitory subunit of[38,39]
phosphopantothenoylprotein phosphatase PPZ1
cysteine decarboxylase
coenzyme A
Peroxiredoxin 1Peroxidase Molecular chaperone[40]
Peroxiredoxin 2PeroxidaseChaperone[40]
Rack1Component of ribosomeScaffold in[41,42]
signal transduction pathways
Succinate dehydrogenaseElectron transport inComponent of TIM22 complex [43,44,45,46]
subunit 3respiratory complex II
Cell Surface
Candida albicans
6-phosphogluconateOxidative decarboxylationBinds components of contact system[47,48,49]
dehydrogenaseof 6-phosphogluconate
Alcohol dehydrogenase 1Alcohol dehydrogenaseBinds plasminogen[50]
Ef-TuElongation factorBinds plasminogen,[48,49,51]
during protein synthesis factor H, HK, FXII, and prekallikrein
EnolaseEnolase, enzymeBinds plasminogen[52]
Fructose bisphosphateFructose bisphosphateBinds plasminogen[50]
aldolase aldolase
GlyceraldehydeGlyceraldehydeBinds plasminogen,[53]
3-phosphate 3-phosphate fibronectin, and laminin
dehydrogenase dehydrogenase
Glutathione reductaseGlutathione reductaseBinds plasma proteins[54,55]
Glycerol 3-phosphateGlycerol 3-phosphateBinds plasminogen[56]
dehydrogenase dehydrogenase
Heat shock protein Ssa2ChaperoneBinds a peptide in saliva [57]
High-affinity glucoseGlucose transporterComplement inhibitor[58]
transporter 1
Integrin-like proteinInhibits opsonizationAdhesin[59,60]
and phagocytosis
Peroxisomal catalasePeroxisomal catalasePlasminogen binding[50]
Phosphoglycerate kinasePhosphoglycerate kinase Plasminogen binding[50]
PhosphoglyceromutasePhosphoglyceromutaseBinds plasminogen, vitronectin, and fibronectin [50,61]
Phosphotransferase Transports andBinds to plasma proteins[48,62]
phosphorylates sugars
Superoxide dismutase 3Removes superoxideBinds contact system proteins[48,49]
radicals HK, FXII, and prekallikrein
Thiol-specific antioxidantThiol-specific antioxidantBinds plasminogen[50]
protein protein
Transcription elongationTranscription elongationBinds plasminogen[50]
factor factor
Triosephosphate IsomeraseBinds extracellular matrix proteins[61]
isomerase
Heat shock protein Ssa2ChaperoneBinds HTN3/histatin 5[57]
Candida glabrata
Triosephosphate IsomeraseBinds extracellular matrix proteins[61]
isomerase
Candida tropicalis
Enolase EnolaseBinds vitronectin, fibronectin,[63,64]
and plasminogen
Fructose-bisphosphate AldolaseBinds fibronectin,[64]
aldolase vitronectin, and laminin
Cryptococcus neoformans
Glucose 6-phosphateGlucose 6-phosphateBinds plasminogen[65]
isomerase isomerase
Nitric oxide dioxygenase Nitric oxide dioxygenaseBinds plasminogen[65]
Phosphoglycerate kinasePhosphoglycerate kinaseBinds plasminogen[65]
Pyruvate kinasePyruvate kinaseBinds plasminogen[65]
Subunits of ATP synthaseATP synthesisBinds plasminogen[65]
Histoplasma capsulatum
Hsp60 chaperoneChaperoneAdhesin[66]
Paracoccidioides brasiliensis
Enolase EnolaseBinds fibronectin [67]
GlyceraldehydeGlycerol 3-phosphateBinds fibronectin, laminin,[68]
3-phosphate dehydrogenaseand type I collagen
dehydrogenase
Triose phosphateTriose phosphateAdhesin[69]
isomerase isomerase
Sporothrix schenckii
GroEL/Hsp60 ChaperoneBinds laminin, elastin,[70]
fibrinogen, and fibronectin
Talaromyces marneffei
GlyceraldehydeGlyceraldehydeAdhesin[71]
3-phosphate 3-phosphate
dehydrogenase dehydrogenase
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Curtis, N.J.; Patel, K.J.; Rizwan, A.; Jeffery, C.J. Moonlighting Proteins: Diverse Functions Found in Fungi. J. Fungi 2023, 9, 1107. https://doi.org/10.3390/jof9111107

AMA Style

Curtis NJ, Patel KJ, Rizwan A, Jeffery CJ. Moonlighting Proteins: Diverse Functions Found in Fungi. Journal of Fungi. 2023; 9(11):1107. https://doi.org/10.3390/jof9111107

Chicago/Turabian Style

Curtis, Nicole J., Krupa J. Patel, Amina Rizwan, and Constance J. Jeffery. 2023. "Moonlighting Proteins: Diverse Functions Found in Fungi" Journal of Fungi 9, no. 11: 1107. https://doi.org/10.3390/jof9111107

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