Next Article in Journal
Molecular Diagnosis of Two Major Implantation Mycoses: Chromoblastomycosis and Sporotrichosis
Next Article in Special Issue
Mycoviromic Analysis Unveils Complex Virus Composition in a Hypovirulent Strain of Sclerotinia sclerotiorum
Previous Article in Journal
Taking a Closer Look: Clinical and Histopathological Characteristics of Culture-Positive versus Culture-Negative Pulmonary Mucormycosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adaptive Response of Saccharomyces Hosts to Totiviridae L-A dsRNA Viruses Is Achieved through Intrinsically Balanced Action of Targeted Transcription Factors

by
Bazilė Ravoitytė
1,*,
Juliana Lukša
1,2,
Ralf Erik Wellinger
3,
Saulius Serva
2,4 and
Elena Servienė
1,4,*
1
Laboratory of Genetics, Institute of Botany, Nature Research Centre, Akademijos Str. 2, 08412 Vilnius, Lithuania
2
Department of Biochemistry and Molecular Biology, Institute of Biosciences, Vilnius University, Saulėtekio Al. 7, 10257 Vilnius, Lithuania
3
Genome Stability Department, Andalusian Center for Molecular Biology and Regenerative Medicine CABIMER, Universidad de Sevilla-CSIC, Avda. Américo Vespucio 24, 41092 Seville, Spain
4
Department of Chemistry and Bioengineering, Vilnius Gediminas Technical University, Saulėtekio Al. 11, 10223 Vilnius, Lithuania
*
Authors to whom correspondence should be addressed.
J. Fungi 2022, 8(4), 381; https://doi.org/10.3390/jof8040381
Submission received: 21 March 2022 / Revised: 5 April 2022 / Accepted: 7 April 2022 / Published: 9 April 2022
(This article belongs to the Special Issue Mycoviruses: Emerging Investigations on Virus-Fungal Host Interaction)

Abstract

:
Totiviridae L-A virus is a widespread yeast dsRNA virus. The persistence of the L-A virus alone appears to be symptomless, but the concomitant presence of a satellite M virus provides a killer trait for the host cell. The presence of L-A dsRNA is common in laboratory, industrial, and wild yeasts, but little is known about the impact of the L-A virus on the host’s gene expression. In this work, based on high-throughput RNA sequencing data analysis, the impact of the L-A virus on whole-genome expression in three different Saccharomyces paradoxus and S. cerevisiae host strains was analyzed. In the presence of the L-A virus, moderate alterations in gene expression were detected, with the least impact on respiration-deficient cells. Remarkably, the transcriptional adaptation of essential genes was limited to genes involved in ribosome biogenesis. Transcriptional responses to L-A maintenance were, nevertheless, similar to those induced upon stress or nutrient availability. Based on these data, we further dissected yeast transcriptional regulators that, in turn, modulate the cellular L-A dsRNA levels. Our findings point to totivirus-driven fine-tuning of the transcriptional landscape in yeasts and uncover signaling pathways employed by dsRNA viruses to establish the stable, yet allegedly profitless, viral infection of fungi.

1. Introduction

Persistent dsRNA viruses in yeasts are usually associated with the killer phenotype of the host cell [1,2,3,4]. Totiviridae L-A virus is an essential component of the dsRNA-based killer system. It encodes the major coat protein (Gag) and RNA-dependent RNA polymerase (Pol), proteins essential for the replication, encapsidation, and maintenance of both the L-A and the M satellite virus, which provide the killer phenotype to the host cell [1]. While dsRNA viruses of fungi often promote hypo- or hypervirulence or a killer phenotype [5,6,7,8], infection with L-A alone seems to be symptomless in yeast [9]. It appears as a result of long-lasting coevolution, as evidenced by the enormous quantity of 4.6 kb-sized L-A dsRNA, which is comparable to that of cellular rRNAs [10].
Several variants of Totiviridae L-A have been described in closely related Saccharomyces cerevisiae and S. paradoxus species [2,3,11,12,13,14,15], and the list of known L-A virus hosts is further expanding [16]. Despite the high average identity of nucleotide sequences, L-A viruses show complex relationships of compatibility with M satellites [14,15,17,18]. The L-A virus seems to be less challenging for the host than satellite M, since only a few genes were previously described to be crucial for L-A maintenance [17]. The factors and mechanisms contributing to L-A virus maintenance are less studied, but L-A dsRNA tends to be more resilient to elimination than M dsRNA [19]. A recent study showed that the overproduction of trimmed L-A capsid protein results in the complete elimination of the virus genome, and does not depend on the L-A type, suggesting similar mechanisms of L-A maintenance in non-killer yeasts [18].
At least several host genes and metabolic conditions are linked to the replication and maintenance of the L-A virus. L-A replication has been reconstructed in vitro and is uniformly believed to occur in cytoplasm via a conservative transcription mechanism [17,20]. N-terminal Gag acetylation by NatC N-acetyltransferase, encoded by the MAK3, MAK10, and MAK31 genes, and other yet-unknown host factors are required for the assembly of L-A virus-like particles [17,21,22,23]. These undetermined factors are likely essential for host viability or substitute one another, thereby complicating their identification. The maintenance of yeast dsRNA viruses is very sensitive to changes in ribosome structure; thus, killer yeasts serve as a convenient model to study the function of ribosomal proteins [24]. Ribosomes are important for L-A propagation [25], and mutations of RPL3, RPD3, SIN3, SAP30, RPL41A, and RPL41B elevate the efficiency of programmed −1 ribosomal frameshifting and often result in a decrease or even in a loss of viral dsRNA(s), mostly affecting satellite M [24,26,27]. A lower L-A dsRNA copy number and the amount of Gag protein were observed in cells with disrupted RPL4A, RPL4B, MAK11, and MAK16 genes, which are involved in the biosynthesis of 60S ribosomal subunits [25]. It was suggested that 60S ribosomal subunits are required for the sufficient translation of L-A capsid proteins; however, there were also exceptions, likely related to the differences in the L-A mRNA sequences between L-A variants [25]. Respiration-deficient non-killer yeast mutants lacking mitochondrial pore protein accumulate large amounts of virus-like particles and L-A dsRNA during adaptation to growth on glycerol [28]. Ethanol-grown yeast cells contain higher levels of viral dsRNA than glucose-grown cells [29], which may be related to the observation that MAK10 is repressed by glucose [30]. It was also proposed that the L-A virus and mitochondrial DNA compete for Mak10 protein [30]. Double-stranded RNA viruses were shown to mask distinct chromosomal mutations and, in some instances, they were affected by a variant of the mitochondrial genome, suggesting that non-chromosomal elements can interact to alter the host phenotype [31].
L-A virus propagation is controlled or inhibited by systems found in yeast cells. Nascent L-A mRNA has neither cap nor polyA tails [32]; thus, its 5′-end is prone to degradation by exonuclease Xrn1/Ski1, whereas cytoplasmic exosome degrades mRNA, which lacks a polyA tail in the 3′-end [12,33]. Xrn1 is involved in the regulation of various cellular processes, including meiosis [34], filamentous growth [35], RNA turnover [36], control of telomere length [37], respiration [38], and autophagy [39]. The action of the exosome is linked to the SKI complex, in which Ski2, Ski3, Ski7, and Ski8 proteins are required for the proper identification of exosome targets (see [40] for review). The disruption of any of these SKI genes enables higher levels of viral dsRNA; in case of the presence of M virus, markedly higher production of killer toxin, leading to the so-called “superkiller” phenotype [33]. To avoid these mechanisms, the virus has developed protection strategies. The cap-snatching mechanism for the translation of L-A proteins was proposed [41]. The ends of L-A positive strands are 5′ diphosphorylated by yet-unknown proteins, even though the triphosphatase and kinase activities have been associated with L-A virus-like particles [42,43]. Diphosphorylation is essential for viral transcription and acts like a self-identity tag of viral RNA [43]. Mitochondrion-related nuclease NUC1 and mitochondrial porin-encoding POR1/2 genes negatively regulate the copy number of L-A dsRNA [28,44,45]. This regulation can be crucial for the survival of sporulating cells [45].
Several studies have highlighted the changes in the L-A virus level in yeast cells upon different metabolic conditions and sporulation [28,44,45]. Recently, host gene transcription in yeasts hosting or cured of dsRNA viruses was examined [46,47,48,49]. Transcriptome studies mainly involve wild-type killer cells as a reference, making the impact of L-A and M viruses overlapping, since killer cells possess two types of dsRNA. While the incidence of killer viruses in the S. cerevisiae and S. paradoxus strains is estimated to be relatively infrequent, L-A dsRNA is nevertheless prevalent in various wild, industrial, clinical, and laboratory yeasts [50]. Biological sensing of the maintenance of the L-A virus in non-killer yeasts, therefore, remains a long-standing challenge in the field of yeast dsRNA viruses.
In this study, we analyzed the transcriptional responses in L-A dsRNA maintaining non-killer S. cerevisiae and S. paradoxus yeasts. We, therefore, strove to analyze the transcriptional pattern of L-A dsRNA maintaining non-killer S. cerevisiae (M437 harboring L-A-lus) and S. paradoxus (AML-15-66 Spo and Pet harboring L-A-66) yeasts based on high- throughput RNA-seq data as compared to dsRNA-free cells, respectively [47,49]. Cured S. cerevisiae cells show the same phenotype as the wild-type cells, except for the absence of the killing ability, while cured S. paradoxus cells gain either respiration-deficient (named Pet) or increased sporulation frequency (named Spo) phenotypes [47,49]. Interestingly, the moderate by intensity, yet broad, transcriptional reprogramming of a host cell upon L-A virus infection is rather yeast strain-specific. Nevertheless, a common hallmark is the transcriptional modulation of genes linked to various stress responses. We found that mutant cells lacking the transcription factors involved in stress response activation are prone to a significant increase or decrease in the L-A dsRNA levels. Thus, our findings shed new light on the host factors beyond the precious balance of L-A maintenance.

2. Materials and Methods

2.1. Gene Expression Data

Processed transcriptomics data were obtained from the Gene Expression Omnibus (GEO) repository under accession numbers GSE100290 [47] and GSE153308 [49]. From the GSE100290 dataset, providing gene expression data in cells derived from S. cerevisiae M437 [kil-K2] [51], three replicates of cells with L-A-lus virus and three virus-free samples were collected. From GSE153308 the dataset, providing transcriptomic data of cells derived from S. paradoxus AML-15-66 [kil-K66] [2], two different (altered-sporulation and respiration-deficient) cell types (three replicates of cells with L-A-66 virus and three virus-free samples for each cell type) were analyzed. All datasets exploited the same yeast growth, RNA extraction conditions, and overall experimental design. Yeast cells were grown in standard liquid complete medium for yeast growth YPD at 25 °C with shaking at 250 rpm. The total RNA for high-throughput RNA sequencing was extracted from cells in the exponential growth phase (OD600 = 0.5–0.6) using GeneJET RNA Purification Kit (Thermo Fisher, Vilnius, Lithuania), according to the Yeast Total RNA Purification Protocol provided by the manufacturer [47]. Detailed descriptions of the experiments that led to the generation of processed transcriptomics data in the GSE100290 and GSE153308 datasets are provided in the following publications [47,49].

2.2. Identification of Differently Expressed Genes (DEGs)

The mean expression values were used to determine differential expression by implementing the “exact test” of Robinson and Smyth [52] by applying the DGE tool in CLC Genomics Workbench v12.0 (CLC Inc., Aarhus, Denmark; http://www.clcbio.com). Gene expression was considered significant if differently expressed genes (DEGs) had a fold-change of ≥1.5 and FDR-corrected p-value of ≤0.05 [53]. Two group comparisons were made between M437[L−M−] vs. M437[L+M−]; Spo[L−M−] vs. Spo[L+M−]; and Pet[L−M−] vs. Pet[L+M−]. Descriptions of the gene products of DEGs were obtained using YeastMine (http://yeastmine.yeastgenome.org, accessed on 15 July 2021) data search and retrieval tool [54] (Table S1).
Volcano plots showing the relationship between the p-value and the magnitude of the difference in the fold-change of the samples were counted and visualized in Microsoft Excel (Figure 1).
Venn diagrams comparing the DEGs of different cells (Figure 2) were generated by using an online tool for the calculation and drawing of custom Venn diagrams (http://bioinformatics.psb.ugent.be/webtools/Venn/, accessed on 4 February 2021).

2.3. GO Pathway Enrichment Analysis

To characterize the relationships between genes, Gene Ontology (GO) analysis was performed. GO TermFinder (http://go.princeton.edu, accessed on 7 September 2021) was used to detect the most significant GO terms [55]. Gene products were annotated by the “Biological processes”, “Molecular function”, and “Cellular component” aspects. The obtained results were analyzed, the statistical significance of each annotation was calculated using hypergeometric distribution, and Benjamini and Hochberg false discovery rate correction was conducted. Results with a p-value of <0.05 were represented (Table S1). The fold-enrichment (F.E.) was calculated by comparing the background frequency of genes annotated to the GO term, with the sample frequency representing the number of input genes that fell under the matching GO term.

2.4. Essential Genes

Essential genes were selected based on information provided in the SGD (Saccharomyces Genome Database [56], http://www.yeastgenome.org, accessed on 22 July 2021). In our study, 506 yeast genes were considered as essential. A total of 474 essential genes were selected with reference to documented evidence provided by experimental data from classical genetics, referred to in the SGD. A gene was assumed as being essential for yeast viability if its complete elimination from the genome was reported to provide lethality in at least one S. cerevisiae yeast strain (CEN.PK, D273-10B, Y55, JK9-3d, S288C, SEY6210, Sigma1278b, SK1, W303, or others). In addition, 32 genes performing essential functions according to the description provided by SGD were also discussed as essential. The systematic gene names of essential genes are marked with different colors in the lists of DEGs in Table S1.

2.5. Determination of Transcriptional Regulators

Transcription factors (TF) linked to the transcription of DEGs were deducted using the tools available in the YEASTRACT (Yeast Search for Transcriptional Regulators AND Consensus Tracking, http://www.yeastract.com, accessed on 17 November 2021) repository [57]. The activity of each TF was predicted by the number of documented targets in the corresponding datasets considering only documented direct regulatory associations based on expression evidence. Lists of upregulated genes were only searched for the targets of transcriptional activators, while those of downregulated genes were only searched for the targets of transcriptional repressors. DEGs were ranked based on the percentage of genes in the list and in the YEASTRACT database regulated by each TF. The p-value was calculated using the hypergeometric test. A full list of TFs is provided in Table S2, where TFs are ranked from the highest to the lowest-confidence-level regulatory associations.

2.6. Strains and Media

To evaluate the effects of gene transcription factors on the L-A dsRNA levels, Saccharomyces cerevisiae strains (BY4741 background, MATa; his3Δ1; leu2Δ0; met15Δ0; and ura3Δ0) were employed. To study the effects of single transcription factor deletions, strains from the S. cerevisiae nonessential gene deletion collection (YKO) with single ORFs replaced by the KanMX4 module were tested. These strains were purchased from Thermo Scientific Molecular Biology (Lafayette, CO, USA). Yeast knockout library (YKO) strains were grown in standard complete medium for yeast growth (YPD) (1% yeast extract, 2% peptone, 2% dextrose; 2% agar was added for solid medium) for 16 h at 25 °C with shaking at 250 rpm.
For transcription factor overexpression experiments, strains from Yeast GST Collection were used (Dharmacon, Lafayette, CO, USA). The overexpression of plasmid-encoded ORFs was under the control of the GAL1/10 promoter. These strains were grown on SD medium (0.67% yeast nitrogen base, L-leucine (60 µg/mL), L-methionine (10 µg/mL), and L-histidine (10 µg/mL), with 2% agar for solid medium) supplemented with 2% glucose for recovery. To induce the overexpression of selected transcription factors, freshly grown cells were grown in liquid SD medium with the required amino acids and 2% galactose (instead of glucose) for 16 h at 25 °C with shaking at 250 rpm.

2.7. Total RNA Extraction and Evaluation of L-A dsRNA Content Changes

Yeast mutant strains with single deletions of transcription factors (except for Met4p, Cyc8p, and Rsc30p) with the highest confidence level (marked in green, Table S2) were subjected to total RNA isolation. Only strains overexpressing selected transcription factors that were significantly altering the L-A dsRNA levels upon deletion were analyzed.
The total RNA was extracted as described in a previous study [47]. The mutant and wild-type strains were grown in YPD medium, whereas strains with plasmid-encoded genes were grown in SD medium with the required amino acids and 2% galactose.
The extracted total RNA was analyzed by 1% agarose gel electrophoresis. The dsRNA amount of L-A virus for each sample was normalized by the 18S rRNA levels and compared to the values of the control samples that were grown in the same medium. Only the strains with the highest difference in the L-A dsRNA levels were considered to be significant.

3. Results

3.1. Transcriptional Response in L-A dsRNA Maintaining S. cerevisiae and S. paradoxus Cells

To better understand the role of L-A dsRNA virus for a host strain, we compared the changes in gene expression between S. cerevisiae M437 and S. paradoxus AML-15-66 cells possessing solely L-A dsRNA in relation to the corresponding dsRNA-free cells (see Table S1 for a full list of DEGs). In addition to focusing on different L-A viruses found in these yeasts (L-A-lus in M437 and L-A-66 in AML-15-66), distinct states of metabolism were addressed: S. cerevisiae M437 (harboring single L-A-lus virus; thus, named M437[L]) features regular glucose-dependent metabolism, while S. paradoxus AML-15-66 cells (bearing solely L-A-66 are named Spo[L] and Pet[L]) indicate altered sporulation and respiration-deficient (petite) phenotypes, as described in the previous study [49].
The presence of L-A dsRNA leads to moderate transcriptional alterations in the examined strains. Volcano plots (Figure 1) represent the confidence levels and fold-changes of DEGs in M437[L], Spo[L], and Pet[L] cells. Approximately 88% of the DEGs in S. cerevisiae M437[L] and 79% of those in S. paradoxus Spo[L] cells were altered up to three times. The expression of 79% of the downregulated genes in S. paradoxus Pet[L] cells was also altered less than three-fold, while 78% of the positively regulated genes were upregulated up to two-fold (Table S1).
Among the most upregulated genes in M437[L] cells, genes encoding mannoproteins related to iron transport (FIT1, -2, -3), other proteins related to iron (LSO1, SIT1), and putative proteins of unknown functions (YHR175W-A, YDR003W-A, and YLR154C-H) (Figure 1 and Table S1) were identified. The most downregulated genes in M437[L] cells were those encoding the cell wall proteins (CWP1, PIR3), protein kinase (KDX1), and proteins involved in cell wall and membrane biosynthesis (OPI3, YMR084W, YMR085W, and INO1).
The genes most upregulated in the Spo[L] cells were AAC3 (encoding mitochondrial ADP/ATP translocator), ASN1 (coding for asparagine synthetase), and those related to ribosome biogenesis (IPI1, CGR1, and UTP18). The PIR3 (encoding cell wall protein), GPX1 (coding for glutathione peroxidase), and SIS2 (encoding a negative regulatory subunit of protein phosphatase 1) genes were the most downregulated in Spo[L] cells (Figure 1 and Table S1).
The most upregulated genes in Pet[L] cells were FLO11 (encoding flocculin), MET17, and MET2 (both involved in the metabolism of methionine). The most downregulated genes in Pet[L] cells were related to the biosynthesis of amino acids (e.g., LYS2, -12, ARG1, -4, -7, and LEU2) (Figure 1 and Table S1).
Overall, gene transcription was the most altered (450 genes) in M437[L] cells, less (250 genes) in Spo[L] cells, and the least (37 genes) in Pet[L] cells (Figure 2 and Table S1). The numbers of up- and downregulated genes in M437[L] cells were similar, with 238 genes being up- and 212 being downregulated. Approximately two-thirds (183 genes) of the DEGs in Spo[L] cells were upregulated, and about one-third (67 genes) were downregulated. An inverse distribution of the gene number ratio was observed in Pet[L] cells, where about one-third of the DEGs were upregulated (nine genes) and two-thirds were downregulated (28 genes) (Table S1). The gene transcription patterns were highly different between each cell type; there were no DEGs shared between all three strains, neither upregulated nor downregulated (Figure 2). Only LYS12 was differently expressed in all three strains; however, it was upregulated in M437[L] and Spo[L], but downregulated in the Pet[L] cells.
The highest number of mutually regulated DEGs was between M437[L] and Spo[L] cells (Figure 2). The majority of spotted genes were linked to the biosynthesis of ribosomes and amino acids. Of the 48 DEGs shared between the M437[L] and Spo[L] cells, 44 were mutually regulated. Thirty-nine of these genes were upregulated and mostly related to ribosome biogenesis (BUD22, DIP2, NSR1, PUF6, RCL1, RPA135, RRP12, and RRS1) and aspartate-family amino acids (AAT1, ACO2, ASN1, IRC7, and ILV1). In M437[L] and Pet[L] cells only, the expression of the MIG2 gene, encoding a protein involved in glucose-dependent gene repression [58], was upregulated. On the other hand, the expression of ZRT1 and MET17 genes, which are linked to zinc uptake [59] and the biosynthesis of sulfur amino acids [60], respectively, were mutually upregulated in Spo[L] and Pet[L] cells.
There was little similarity between the profiles of downregulated genes in the tested cells (Figure 2). The OPI3, PIR3, EHT1, GPX1, and RCN1 genes were negatively regulated in both M437[L] and Spo[L] cells, and most of these genes are linked to the metabolism of phospholipids. Only the SSA4 chaperone-encoding gene was mutually downregulated in S. paradoxus Spo[L] and Pet[L] cells. MAK21, involved in ribosome biogenesis and dsRNA maintenance, was upregulated in M437[L] cells.

3.2. Gene Ontology Analysis of DEGs in Solely L-A dsRNA Maintaining Cells

GO analysis of DEGs revealed similarities and differences between S. cerevisiae M437[L] and S. paradoxus Spo[L] cells and Pet[L] cells. The analysis of DEGs showed nine mutual GO process terms and one GO function term among the three cell types (Figure 3 and Figure 4, Table S1). Table S1 contains a complete set of GO process, function, and localization analysis results, as well as the calculated fold-enrichment (F.E.) values. The small-molecule, oxoacid, organic, and carboxylic acid metabolic processes were enriched by downregulated genes in M437[L] and upregulated genes in Spo[L] and Pet[L] cells (Table S1). The catalytic activity function was represented by negatively regulated genes in M437[L] and Pet[L] cells and by upregulated genes in Spo[L] cells.
The regulation of genes involved in the metabolism of amino acids was altered in all cell types (Figure 3 and Figure 4, Table S1). Positively regulated genes in M437[L] and Spo[L] cells and negatively regulated genes in Pet[L] cells were related to the biosynthesis and metabolism of the aspartate family and alpha-amino acids (Table S1). Lysine metabolism was represented by upregulated genes in M437[L] cells and by downregulated DEGs in Pet[L] cells (F.E. of 12.5 and 107.0, respectively). Arginine biosynthesis was associated with genes downregulated in Pet[L] and upregulated in Spo[L] cells (F.E. of 89.5 and 13.5, respectively). The metabolism of sulfur amino acids and methionine (F.E. of 7.0 and 8.2, respectively) was upregulated in Spo[L] cells. Genes related to the metabolism of the histidine, aromatic, and glutamine amino acid families (F.E. of 107.0, 50.0, and 27.8, respectively) were downregulated in Pet[L] cells.
RNA-related processes were represented by upregulated genes in M437[L] and Spo[L] cells (Figure 3 and Figure 4, Table S1). DEGs related to RNA processing and metabolism, including ncRNA and rRNA processing, the biogenesis of ribosomes (F.E. of 3.4 and 5.6), and transcription by RNA polymerase I (F.E. of 5.2 and 5.5) were upregulated in the M437[L] and Spo[L] cells. The binding of snoRNA (F.E. of 12.5 and 6.6) was also represented by positively regulated genes in these cells. The activities of RNA helicase (F.E. of 6.3), RNA methyltransferase (F.E. of 5.8), and tRNA modification (F.E. of 5.0) were enriched in M437[L] cells only (Table S1). The binding of mRNA (F.E. of 8.2) and the activity of RNA polymerase III (F.E. of 4.4) were enriched solely in Spo[L] cells. GO Cellular component enrichment analysis showed that many of the genes upregulated in the M347[L] and Spo[L] cells were associated with gene product localization in the nucleus (F.E. of 1.4 and 1.5), nucleolus (F.E. of 4.2 and 6.1), and pre-ribosome (F.E. of 5.6 and 8.9).
Gene ontology analysis revealed that, in M437[L] cells, upregulated genes were linked to ion transport, homeostasis, and transmembrane transport (Figure 3). The transport of cofactors, copper ions, iron ions, and siderophores was enriched (F.E. of 6.8, 11.3, 9.7, and 17.0, respectively) with positively regulated genes in M437[L] cells. Functional GO term analysis uncovered enrichment in the transporter and ferric-chelate reductase activities (F.E. of 2.2 and 17.0). Positively regulated genes in Pet[L] cells were related to homoserine metabolism (F.E. of 111.0), cofactor binding (F.E. of 11.5), and mitochondrial nucleoids (F.E. of 55.5); meanwhile, in Spo[L] cells, similarly to M437[L], upregulated DEGs were primarily involved in various RNA and amino acid processes (Figure 3 and Figure 4, Table S1).
Negatively regulated genes in M437[L] cells were related to the response to chemical stimulus (F.E. of 2.4) and unfolded proteins (F.E. of 7.0), cell wall (F.E. of 3.3), and respiratory electron transport chain (F.E. of 6.3) (Figure 3 and Table S1). Cell wall chitin and polysaccharides, UDP-N-acetylglucosamine metabolism, and the biosynthesis of phosphatidylcholine and phosphatidylethanolamine were enriched by downregulated genes in M437[L] cells (F.E. of 9.3, 6.6, 19.0, 9.3, and 19.0, accordingly). Downregulated genes in Pet[L] cells were mostly linked to amino acids, while those in Spo[L] cells were associated with the mitochondrial gene expression (F.E. of 5.1) and translation (F.E. of 6.0) (Figure 4). Genes related to mitochondria are represented by enriched structural constituents of the ribosome (F.E. of 13.4), mitochondrial ribosome (F.E. of 9.15), and mitochondrial intermembrane space (F.E. of 10.0) GO terms.

3.3. Differently Expressed Essential Genes

We found that the majority of DEGs altered upon L-A infection are not essential for yeast viability. Out of 506 essential genes, 88 were differently expressed in at least one type of solely L-A-maintaining cells. The transcription of 49, 46, and 1 essential gene was altered in the M437[L], Spo[L], and Pet[L] cells, respectively (Table S1). More than half of these genes are related to the biosynthesis of ribosomes (DBP-, NOP-, and RPA- genes) and tRNA transcription (RPC- genes, encoding RNA polymerase III subunits).
The majority of differently expressed essential genes were upregulated in the Spo[L] and M437[L] cells. There were seven shared essential genes of altered transcription between M437[L] and Spo[L] cells, and all were upregulated. Only CAB1 was involved in coenzyme A biosynthesis, while others (DIP2, RCL1 RPA135, RRS1, SPB1, and UTP18) were related to ribosomes. GGC1, which is essential for mitochondrial genome maintenance and important for mitochondrial iron transport, was upregulated in M437[L] and downregulated in Pet[L] cells. Upregulated essential genes in M437[L] cells were also related to iron–sulfur proteins (CFD1 and YAH1) and the nuclear envelope (BRL1), whereas those in Spo[L] cells were involved in ergosterol biosynthesis (ERG1, -11), the composition of RNA polymerases (RPA-, RPB8, RPO26, RPC- genes), pentose phosphate pathway (RKI1), and tRNA synthesis (WRS1). RLI1, encoding an essential iron–sulfur protein required for ribosome biogenesis, was upregulated in Spo[L] cells. MAK5 was upregulated in M437[L] cells.
Only four essential genes (HSK3, EAP1, UGP1, and MIA40) were negatively regulated in Spo[L] cells. Downregulated essential genes in M437[L] cells were related to the folding, modification, and transport of proteins (SSC1, HSP60, HSC82, PDI1, VPS17, UFD1, and CAR2), and cell wall biogenesis (DFG5 and PCM1). RNA exonuclease encoding USB1 was the most downregulated essential gene in M437[L] cells.

3.4. Transcription Factors Regulating DEGs in Solely L-A-Infected Cells

To assess the possible role of transcription factors (TFs) in the global transcriptional alterations in L-A virus maintaining cells, we employed YEASTRACT analysis tools [57]. We aimed to identify which TFs could be the key regulators (activators or inhibitors) of gene expression in virus-infected cells. We inspected the documented transcription activators of upregulated genes and inhibitors of downregulated genes in solely L-A-maintaining cells in respect of dsRNA-free cells (Figure 5 and Table S2).
Our analysis revealed that 39 TFs regulated the expression of DEGs at the highest confidence level (Figure 5 and Table S2). Twenty-four TFs were associated with the activation of upregulated genes and 18 TFs were linked with the inhibition of downregulated genes (Figure 5). Out of these, three TFs were involved in both the up- and downregulation of DEGs, depending on the strain (Figure 5). A significant part of the TFs involved in the regulation of DEGs was associated with the response to various stresses (e.g., Yap1, Pdr3, Rtg1, Rtg2, Yrr1, Stb5, Mga2, Cst6, Sfp1, Eds1, Cup2, Wtm2, and Rsc30), response to nutrients (e.g., Mss11, Gat1, Gcr1, Mal33, Nrg2, and Mig2), and related to amino acids (e.g., Leu3, Stp1, Met4, and Lys14).
Only several TFs were related with different expressions of a large part of the DEGs. The expression of 53% of the upregulated DEGs in M437[L] cells can be induced by Gat1, and the highest number of downregulated genes (56%) can be associated with Bas1-mediated repression. The transcription of the majority of positively regulated genes of Spo[L] cells is related to the action of Cst6 (62%), Gat1 (61%), and Sfp1 (59%) TFs. The expression of the highest percentage (45%) of downregulated genes in Spo[L] can be repressed by Mal33. Up to 100% of the DEGs in Pet[L] cells can be differently expressed by TFs Yap1 (100% of upregulated genes), Pdr3 (100% of upregulated genes), and Yrr1 (89% of up- and 86% of downregulated genes).
Among the six datasets (up- and downregulated genes of each cell type) of DEGs in M437[L], Spo[L], and Pet[L] cells, the gene expression pattern in M437[L] and Spo[L] cells seemed to be the most similar in terms of both shared genes (Figure 2 and Table S1) and related TFs (Figure 5 and Table S2). TFs Stb5 and Gat1 may activate the expression of a part of the upregulated genes, while Cup2 may repress the expression of a fraction of the downregulated genes in M437[L] and Spo[L] and cells (Figure 5 and Table S2). TF Yrr1 may be important for controlling the downregulation of 49% of genes in M437[L] cells, and for the vast majority of DEGs in Pet[L] cells. Sut1 may have an impact on the expression of 25% of up- and 34% of downregulated DEGs in Spo[L] cells. Mga2 protein could activate the transcription of 36% of the upregulated genes in Spo[L] and repress 41% of the downregulated genes in M437[L] cells.
The expression of three genes encoding the transcription factors Bas1, Mig2, and Rsc30 was upregulated in M437[L] cells, and MIG2 was also upregulated in Pet[L] cells. Bas1 is involved in the repression of 56% of the downregulated genes in M437[L] cells (Figure 5 and Table S2). TF Rsc30 regulates 9% of the upregulated genes in M437[L] cells. Mig2 is associated with the activation of about 12% of the genes upregulated in M437[L] cells. This observation suggests that TFs may be involved in the regulation of L-A virus-related gene expression alterations and may be important for virus maintenance.

3.5. Role of Transcription Factors on Intracellular L-A dsRNA Content

Strains based on S. cerevisiae M437 and S. paradoxus AML-15-66 are prototrophic diploid yeasts; thus, genetic manipulations are challenging and labor-intensive. To determine the impact of TFs on the actual intracellular levels of the L-A virus genome, we employed laboratory S. cerevisiae yeast strains of BY4741 background. The parental S. cerevisiae BY4741 strain is haploid, natively harbors L-A-1 dsRNA, and is devoid of M satellite; thus, it is a non-killer strain. The selection of this genetic background enabled us to verify the relevance of observed transcriptional regulation patterns linked to the L-A virus in wild yeasts.
We selected the TFs listed in Figure 5 and tested how the elimination of each TF would affect the intracellular L-A dsRNA content in the BY4741 background. The single deletion of 13 TFs significantly altered L-A dsRNA levels (Figure 6A). The removal of two transcription activators (Δgcr1 and Δpdr3) resulted in increased L-A levels, while the elimination of the other three (Δmss11, Δsfp1, and Δylr278c) led to decreased amounts of L-A dsRNA. The deletion of one out of four transcription inhibitors (Δopi1, Δrph1, Δtog1, or Δume6) was linked to an increased L-A dsRNA content, while the elimination of the other four (Δeds1, Δleu3, Δmal33, or Δnrg2) led to a decreased L-A dsRNA content (Figure 6A). The removal of one of the other tested TFs did not lead to an observable change in the L-A dsRNA copy number (not shown); however, their role should not be neglected. Other TFs may have a more significant impact under other genetic backgrounds or growth conditions.
The overexpression effect of TFs, involved in significant L-A dsRNA content changes observed upon TF deletion, was analyzed. The overexpression of seven TFs (EDS1, LEU3, MAL33, NRG2, OPI1, TOG1, and UME6) led to higher amounts of L-A dsRNA (Figure 6B), whereas, in cells overexpressing MSS11, PDR3, or SFP1 genes, the L-A dsRNA levels were similar to the wild type (Figure 6B). There was no decrease in L-A dsRNA in cells overexpressing any of the tested TFs, which may be related to galactose metabolism, since it was observed that more of L-A dsRNAs accumulate in cells grown on a nonfermentable carbon source [28,29].
Since L-A virus propagation requires the acetylation of the Gag protein [17,21,22], the impact of several genes linked to the regulation of the intracellular acetyl-CoA levels was analyzed (Figure 6C). In concordance with acetyl-CoA homeostasis being important for L-A virus propagation, the L-A dsRNA level was found to be decreased in Δcit1, Δcat2, and Δpcd1, but not in Δpda1, Δach1, Δacs1, or Δsnf1 strains.

4. Discussion

4.1. Gene Expression in Yeast Cells Maintaining Totiviridae L-A Virus

The adaptive modulation of host gene expression by the L-A virus likely evolved as a useful strategy for virus persistence and might result from long-term co-evolution. Consistent with previous findings [46,47,49], our data indicate that L-A dsRNA has a rather minor role in the host gene expression. Of the 450 DEGs in M437[L] cells, 328 were also differently expressed alike upon the elimination of solely L-A or both L-A and M dsRNAs, when compared to the wild-type killer cells [47], including the upregulation of MAK10, the essential gene for dsRNA replication [22]. Only two new DEGs (MET2 and SSA4) were found in S. paradoxus AML-15-66 Pet[L] cells, since 35 genes were also differently expressed in cured Pet cells with respect to the wild-type killer cells [49]. Twenty-seven genes were also differently expressed in cured Pet cells, three genes (FLO11, MET17, and HSP26) were solely expressed in dsRNA-free Pet cells, and five genes (MUP1, ACO1, IDH1, HIS4, and LYS20) were only expressed in Pet cells possessing solely L-A dsRNA. Compared to previous observations, we conclude that it is not the helper L-A, but satellite M, virus that has a significant influence on the host transcriptome [47,49]. The satellite M virus not only employs L-A capsids for its genome replication, but it also utilizes the protein synthesis, maturation, modification, and secretion pathways of the yeast cell to promote killer toxin production and the maintenance of self-immunity [1,61,62,63]. Accordingly, killer toxin production is also associated with transcriptional responses linked to increased host cell stress tolerance [61].
Gene expression in Pet[L] cells seems to be the most distinct based on the number of genes affected and the relative alteration of gene expression. However, it is likely that this difference is also related to the respiration-deficient phenotype of Pet[L] cells [49]. Petite cells exhibit many physiological and metabolic differences in relation to the wild-type cells, including those with a small colony size, increased cell adhesion, low acetate catabolism, and inability to utilize nonfermentable carbon sources [64,65]. Notably, this phenotype might be advantageous, since it has been shown to promote heat resistance, to confer pleiotropic drug resistance, and even to prolong lifespan [64,66,67].
Interaction between the essential genes and L-A virus might be important for virus persistence. The essential role of mitochondria and iron–sulfur clusters in ribosome biogenesis [68] can be related to the strong upregulation of FIT genes, which is important for iron import in M437[L] cells. MAK5 [69], encoding for an essential nucleolar protein required for dsRNA maintenance, was also upregulated in M437[L] cells. This close relationship between the essential host genes and the virus may ensure that virus is transmitted to a healthy and viable host. It remains unclear why genes involved in mitochondrial gene expression and translation are downregulated in Spo[L] cells, although mitochondrial functions and proteins have been related with dsRNA maintenance. L-A virus replication is believed to occur in the cytoplasm; however, there is evidence that replicases of other RNA viruses colocalize with mitochondrial membranes; thus, mitochondria might also be related to the replication of the Totiviridae L-A virus [70]. The L-A virus-hosting Pet[L] strain is characterized by a reduction in genes prone to transcriptional alterations. This finding could provide further clues regarding the importance of yeast respiration in viral dsRNA maintenance. It would be interesting to determine if, at some point, the L-A virus becomes essential for yeast viability by driving the transcription of essential host genes. It remains to be determined if L-A viruses could help yeast to adapt to metabolic needs or other kinds of stress, or even provide an epigenetic memory for adaptive responses.
L-A-related transcriptional alterations are somewhat similar to those of other dsRNA mycoviruses. The infection of dimorphic fungi Talaromyces marneffei with dsRNA partitivirus TmPV1 resulted in the altered transcription of genes involved in the regulation of the transcription, transport, and metabolism of amino acids, lipids, and polysaccharides [71]. More viral genetic material was found in yeast than in the mycelial phase of T. marneffei and dsRNA enhanced the virulence of the fungi [71]. TmPV1 caused the suppression of RNA-interference genes in T. marneffei, suggesting that similar transcriptional reprogramming could have been performed by ancestral L-A dsRNA and contribute to the loss of RNA interference in many Saccharomyces and other yeast species [71,72]. Transcriptomic responses in totivirus-infected Malassezia yeasts were mostly related to ribosome biogenesis and other cellular processes [73,74]. The transcription of ribosomal RNA genes was also altered in Sclerotinia sclerotiorum fungus upon infection with Hypoviridae SsHV2-L dsRNA virus [75]. The importance of ribosomes and amino acids for the propagation of L-A dsRNA and its proteins has been documented [17,25]. It would be interesting to determine whether infection with L-A could contribute to yeast pathogenicity in symbiotic relationships with other living organisms, similarly to Totiviridae Leishmania virus 1 of the human pathogen Leishmania braziliensis, which is associated with the troublesome curing of leishmaniasis [76,77]. Our findings suggest that different host metabolic conditions can lead to distinct transcriptional responses in virus-infected and dsRNA-free cells. Targeting similar metabolic functions in other viral dsRNA-infected organisms may provoke a switch from harmless infection to pathogenesis [78]. It has already been demonstrated that virus-infected and virus-free yeasts show differences in their transcriptional responses to antiviral drugs [48].
The gene expression patterns induced by the L-A virus appear to be similar to those induced by various nutrient- and stress-related responses, and they are strongly linked to the host cell’s characteristics. The actual upregulation of genes encoding Bas1, Mig2, and Rsc30 TFs is consistent with the gene expression pattern of M437[L] cells. Mig2 and Rsc30 may be involved in upregulation and act as either direct or indirect activators of the expression of upregulated genes, whereas Bas1 may act as a transcriptional inhibitor and contribute to the observed downregulation of DEGs in cells infected with L-A virus. Bas1 is considered to control the response to a variety of stresses [79]. The Mig2 protein is involved in glucose-induced gene repression [58], and studies suggest that Mig2 plays a prominent role in the control of the yeast metabolic cycle and is linked to the mitotic cell cycle, metabolism of carbohydrates and amino acids, TCA cycle, endonuclease activity, endocytosis [80], and filamentous growth [81]. Rsc30 regulates the expression of ribosomal genes and the cell wall stress response [82]. The fact that we did not observe a significant overlap in TFs involved in the expression of the majority of DEGs in the different yeast species analyzed points to a complex and unique regulatory, transcriptional network for each yeast species [83], which is correspondingly adapted to the co-evolution of the virus and its host cell.

4.2. Interconnection between Transcription Factors and L-A dsRNA Levels

We hypothesize that some DEGs in L-A-maintaining cells are either virus-favoring or virus-inhibiting, and transcriptional regulation is maintained mainly by cellular TFs (Figure 7A). We assumed that higher expression of virus-favoring genes may increase the amount of intracellular L-A dsRNAs, while the expression of virus-inhibiting genes could result in decreased virus propagation. We propose that L-A virus’ persistence is accompanied by increased expression of genes favoring and genes inhibiting L-A maintenance (Figure 7A). Numerous TFs are involved in this complex regulatory network. The virus might act towards the activation of genes required for its replication (e.g., MAK3, MAK10, and MAK31 [17,21,22]), whereas the host may induce the expression of genes involved in virus suppression (e.g., XRN1/SKI1, SKI2, SKI3, SKI7, SKI8, NUC1, and POR1/2 [28,33,44,45]) to control virus propagation and vice versa. Similarly, the virus may contribute to the repression of the transcription of genes conflicting with its functioning; similarly, the host might inhibit the expression of genes that aid the virus. The balance of regulation must be met to optimize the functioning of the virus–host system. Although the regulation of gene expression in yeasts is complex and is not limited to TFs (reviewed in [84,85]), in our model, we simplified the regulation of gene transcription. We assumed that TFs directly or indirectly regulate the transcription of genes, linked to the dsRNA amount in the L-A virus (Figure 7A). We predicted the possible effect on the L-A dsRNA levels if TF involved in any type of transcriptional regulation is eliminated (ΔTF) or overexpressed (TFOE).
Our data support the hypothesis based on the effects of both elimination and overexpression for the role of at least four TFs, namely Eds1, Leu3, Mal33, and Nrg2. All of these TFs act in a way to significantly increase the amount of L-A virus, likely by lowering the expression of virus-inhibiting genes (corresponding to the yellow TF in Figure 7A). In Reimand et al.’s study [86], Eds1 and Mal33 were associated with the transcriptional repression of POR1, which is known to negatively affect the L-A copy number [28], confirming that our results are consistent with previous findings. Other TFs affecting virus propagation are also associated with the transcriptional regulation of genes linked to the L-A virus levels; however, we did not observe a lower L-A dsRNA amount in cells upregulating any of the tested TFs (Figure 6B). This indicates the eminent importance of carbon metabolism in the regulation of the L-A levels, particularly in the absence of glucose.
Out of the 13 TFs that significantly contributed to the L-A dsRNA levels, nine TFs were linked with glucose metabolism, two TFs were associated with ribosome biogenesis (Rph1, Sfp1, and YLR278C), and one was associated with amino acid biosynthesis (Leu3).
Sfp1 activates the expression of ribosomal protein genes and is responsive to nutrients and stress [87]. Sfp1 could act in all groups of TFs in our model (Figure 7A), based on the published data. Sfp1 was associated with the positive regulation of the expression of the MAK10 gene, which is beneficial for L-A virus maintenance [88] (Figure 7A, orange TF), but also of genes linked to L-A virus repression (Figure 7A, green TF), namely NUC1, POR1, and SKI2 [86,89]. Along this line, Sfp1 is associated with the inhibition of MAK3 expression [86], which is required for L-A propagation (Figure 7A, blue TF), and of POR2, and SKI7 [86,90]—genes linked to L-A virus repression (Figure 7A, yellow TF). Interestingly, Sfp1 regulates its own gene transcription, in addition to other genes coding for TFs involved in significant alterations of the L-A levels, namely Eds1, Nrg2, Pdr3, Rph1, and Gcr1 [86,90,91].
Seven TFs were associated with similar alterations in the L-A dsRNA levels, namely decreasing upon TF deletion, and increasing or exhibiting no change upon TF overexpression (Figure 7), such as Sef1. The elimination or overexpression of NRG2, EDS1, and LEU3, results in decreased or increased levels of L-A dsRNA. Nrg2 mediates glucose repression and a group of stress-response genes [92]. The function of Eds1 is suggested by its paralog Rgt1, a glucose-responsive TF regulating glucose transport [93]; its transcription is not only regulated by Sfp1, but also by Leu3 [86]. Leu3 regulates the transcription of genes involved in the biosynthesis of branched-chain amino acids [94]. Mal33 and Mss11 are both related to the regulation of carbohydrate metabolism [95,96], and their elimination results in a decreased amount of L-A virus (Figure 7A). Mss11 is also involved in cellular signaling in response to nutrient levels [96]. The function of YLR278C is unknown, but it is localized in the nucleus and a certain part of the YLR278C-encoded protein enhances the formation of [URE3] prion, and is also involved in the regulation of the yeast metabolic cycle and chromatin remodeling [80,97].
Six TFs were linked to increased L-A dsRNA levels upon TF elimination or TF overexpression, except for PDR3, the overexpression of which resulted in the wild-type L-A level (Figure 7). Most of these TFs are related to glucose; Tog1 activates gluconeogenesis genes and non-fermentable carbon metabolism [98], Gcr1 activates glycolysis genes [99], and Ume6 is a glucose-dependent repressor [100]. Opi1 is required for normal cardiolipin levels and mitochondrial function [101], whereas Pdr3 responds to mitochondrial protein import stress and is involved in multidrug resistance signaling [102]. Rph1 suppresses the genes of ribosome biogenesis in response to nutrient levels [103].
These observations suggest that TFs, changing the level of the L-A virus, may act via the regulation of ribosome biogenesis, glucose-dependent gene expression, and the acetyl-CoA level (Figure 7B). Rph1 and Sfp1 appear to be the main TFs regulating ribosome biogenesis and the quantity of L-A virus. Consistent with the requirement of ribosomes for the translation and co-translational acetylation of viral proteins, Δrph1 cells lack the repression of ribosome biogenesis and have more L-A dsRNA. However, Δsfp1 lacks activation of ribosome biogenesis genes and leads to a decreased virus level (Figure 6A). Glucose represses MAK10 expression [30]; thus, it may negatively regulate the Gag protein acetylation and assembly of L-A particles. We suggest that acetyl-CoA is the central metabolite that determines not only the efficacy of L-A virus propagation, but also the energetic state of the cell. It is involved in the regulation of gene expression through the acetylation of histones, TFs, and enzymes, and affects cellular metabolism and growth (reviewed in [104]). The generation of acetyl-CoA occurs in the mitochondrial and nucleo-cytosolic compartments of the cell, where the catabolism of glucose, fatty acids, and branched-chain amino acids is a major driver [104].
Based on this idea, we aimed to analyze the L-A virus abundance in cells that have genetic defects related to acetyl-CoA turnover (Figure 6C). Interestingly, we observed a decrease in the amount of L-A viral genome in a Δcit1 mutant strain lacking mitochondrial citrate synthase. This mutant strain is believed to contain a higher level of mitochondrial acetyl-CoA due to its inability to use acetyl-CoA for citrate synthesis [105]. As Gag protein acetylation is believed to occur in cytoplasm, a decrease in the quantity of L-A dsRNA in Δcit1 strain may be linked to the increased acetylation of mitochondrial proteins [105], or to impaired citrate synthesis, as citrate is needed as a metabolite for cytoplasmic acetyl-CoA production. Along this line, lower L-A abundance was detected in Δcat2 strain, consistent with the disruption of acetyl-CoA transport into the cytoplasm in a form of acetyl-carnitine from peroxisomes in this mutant [105]. The elimination of PDC1 resulted in diminished L-A levels, likely due to the lower capacity of β-oxidation in peroxisomes and lower generation of acetyl-CoA therein [106]. The L-A levels in other mutants tested did not change, suggesting that a subunit of the pyruvate dehydrogenase Pda1, acetyl-CoA hydrolase/transferase Ach1, and acetyl-CoA synthetase Acs1 had no significant impact on L-A propagation in glucose medium (Figure 6C). We also tested the role of protein kinase Snf1, which is required for proper glucose-dependent gene expression and cellular acetyl-CoA levels [107,108]. Even though Δsnf1 was reported to have about 70% less acetyl-CoA than the wild-type strain [108], the L-A level in this mutant was not altered. Current methods do not allow for the quantification of acetyl-CoA levels in distinct cellular compartments; thus, it remains a challenging task to determine the relationship between cytosolic acetyl-CoA level and L-A maintenance.
The crystal structure of the yeast NatC complex has been recently resolved, and a high efficiency for L-A Gag N-terminus acetylation was observed, pointing to the evolutionary adaptation of the virus [109]. There is a relatively low number of yeast proteins that are acetylated by the NatC complex in comparison to the other acetyltransferases; however, only three mitochondrial proteins (namely Kgd1, Fum1, and Mrp1) are known to have identical N-termini to the Gag protein and are acetylated in a similar manner [110]. The NatC complex is evolutionarily conserved from yeast to humans, but the genetic background is important for complementation, and differences are believed to be linked to mitochondrial function and stress resistance [111]. In line with acetyl-CoA being a central intermediate of the pathways required to metabolize non-fermentable carbon sources [104], the disruption of the NatC complex results in reduced yeast growth on non-fermentable carbon sources [23,28,111]. N-terminal acetylation is thought to be an irreversible modification, and the alteration of this process is linked to the development of disease in humans [112,113]. The function and substrates of NatC complex subunits need to be clarified.
The regulatory associations between TFs and L-A virus propagation require more investigation. The cell and TFs do not discriminate between genes affecting the L-A levels per se. The same TFs likely control both types of genes and have overlapping functions or even opposite regulation patterns depending on the environmental factors and active cellular metabolism pathways. This insight helps to comprehend why, upon L-A virus infection in one strain, certain genes are upregulated, while they are downregulated in the other. We also observed this contradiction with the TFs Yrr1, Mga2, and Sut1, since they were associated with both the up- and downregulation of DEGs in different strains (Figure 5). Yrr1 is associated with the pleiotropic drug resistance pathway [114]. The Mga2 protein is involved in the response to hypoxia, cobalt ion treatment, and iron ion deprivation [115,116]. Sut1 regulates various processes (sterol uptake [117], hypoxic gene expression [118], and the repression of filamentation-inducing genes [119]) in response to pheromones [120]. The elimination of the majority of tested TFs did not significantly change the L-A dsRNA level, probably due to the long-lasting coevolution of the host and virus establishing the self-compensating transcriptional landscape.

5. Conclusions

The persistence of L-A dsRNA leads to an adaptive transcriptional response in S. cerevisiae and S. paradoxus hosts. The least impact of the L-A virus on the host gene expression was observed in the respiration-deficient (S. paradoxus Pet[L]) strain, while the most changes were induced in S. cerevisiae (M437[L]) cells. The majority of DEGs were related to the biosynthesis of amino acids and ribosomes, including a part of the essential host genes. Altered transcription patterns were similar to those induced upon various nutrient and stress responses and dsRNA-virus infections in fungi. We have revealed the role of several yeast transcription factors predicted to regulate gene expression in L-A-maintaining cells based on experimental data and our model. The elimination of most of the tested transcription factors did not result in a significant change in the L-A genome levels; however, Δgcr1, Δopi1, Δpdr3, Δrph1, Δtog1, or Δume6 resulted in an increased L-A dsRNA amount, while Δeds1, Δleu3, Δmal33, or Δnrg2 resulted in a decreased L-A dsRNA amount in S. cerevisiae cells. The overexpression of genes encoding the transcription factors Eds1, Leu3, Mal33, Nrg2, Opi1, Tog1, and Ume6 led to increased L-A genome levels in S. cerevisiae cells grown in galactose medium. The elimination of CIT1, CAT2, and PCD1 genes resulted in reduced intracellular L-A viral genome levels, further indicating an important role of host cell acetyl-CoA-related metabolism viral maintenance. The yeast respiration mode is yet another factor regulating the transcriptional response to the presence of L-A virus in Saccharomyces paradoxus. We observed significant species- and metabolism-oriented specificity in alterations of the transcriptional landscape stemming from the presence of the L-A virus. Taken together, our findings further advance our understanding of the adaptive mechanism employed by widespread Totiviridae viruses to promote viral propagation in yeast cells.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jof8040381/s1, Table S1: Differently expressed genes and GO terms of altered transcription genes in M437[L], Spo[L], and Pet[L] cells. Essential genes are indicated by color; Table S2: Transcription factors affecting the expression of differently expressed genes in M437[L], Spo[L], and Pet[L] cells.

Author Contributions

Conceptualization, B.R., J.L. and E.S.; methodology, B.R. and J.L.; software, J.L.; validation, B.R. and J.L.; formal analysis, B.R., J.L., R.E.W. and E.S.; investigation, B.R. and J.L.; resources, E.S.; data curation, J.L. and B.R.; writing—original draft preparation, B.R.; writing—review and editing, B.R., E.S., J.L., S.S. and R.E.W.; visualization, J.L.; supervision, E.S.; funding acquisition, J.L. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Social Fund under the No. 09.3.3-LMT-K-712-19-0157 “Development of Competences of Scientists, other Researchers and Students through Practical Research Activities” measure.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

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 results.

References

  1. Schmitt, M.J.; Breinig, F. Yeast Viral Killer Toxins: Lethality and Self-Protection. Nat. Rev. Microbiol. 2006, 4, 212–221. [Google Scholar] [CrossRef] [PubMed]
  2. Vepštaitė-Monstavičė, I.; Lukša, J.; Konovalovas, A.; Ežerskytė, D.; Stanevičienė, R.; Strazdaitė-Žielienė, Ž.; Serva, S.; Servienė, E. Saccharomyces paradoxus K66 Killer System Evidences Expanded Assortment of Helper and Satellite Viruses. Viruses 2018, 10, 564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rodríguez-Cousiño, N.; Gomez, P.; Esteban, R. Variation and Distribution of L-A Helper Totiviruses in Saccharomyces Sensu Stricto Yeasts Producing Different Killer Toxins. Toxins 2017, 9, 313. [Google Scholar] [CrossRef] [PubMed]
  4. Wickner, R.B. “Killer Character” of Saccharomyces cerevisiae: Curing by Growth at Elevated Temperature. J. Bacteriol. 1974, 117, 1356–1357. [Google Scholar] [CrossRef] [Green Version]
  5. Nuss, D.L. Hypovirulence: Mycoviruses at the Fungal-Plant Interface. Nat. Rev. Microbiol. 2005, 3, 632–642. [Google Scholar] [CrossRef]
  6. Ahn, I.P.; Lee, Y.H. A Viral Double-Stranded RNA Upregulates the Fungal Virulence of Nectria Radicicola. Mol. Plant. Microbe Interact. 2001, 14, 496–507. [Google Scholar] [CrossRef] [Green Version]
  7. Ejmal, M.A.; Holland, D.J.; MacDiarmid, R.M.; Pearson, M.N. A Novel Chrysovirus from a Clinical Isolate of Aspergillus thermomutatus Affects Sporulation. PLoS ONE 2018, 13, e0209443. [Google Scholar] [CrossRef]
  8. Mannazzu, I.; Domizio, P.; Carboni, G.; Zara, S.; Zara, G.; Comitini, F.; Budroni, M.; Ciani, M. Yeast Killer Toxins: From Ecological Significance to Application. Crit. Rev. Biotechnol. 2019, 39, 603–617. [Google Scholar] [CrossRef]
  9. Ghabrial, S.A. Origin, Adaptation and Evolutionary Pathways of Fungal Viruses. Virus Genes 1998, 16, 119–131. [Google Scholar] [CrossRef]
  10. Ball, S.G.; Tirtiaux, C.; Wickner, R.B. Genetic Control of L-A and L-(BC) DsRNA Copy Number in Killer Systems of Saccharomyces cerevisiae. Genetics 1984, 107, 199–217. [Google Scholar] [CrossRef]
  11. Icho, T.; Wickner, R.B. The Double-Stranded RNA Genome of Yeast Virus L-A Encodes Its Own Putative RNA Polymerase by Fusing Two Open Reading Frames. J. Biol. Chem. 1989, 264, 6716–6723. [Google Scholar] [CrossRef]
  12. Rowley, P.A.; Ho, B.; Bushong, S.; Johnson, A.; Sawyer, S.L. XRN1 Is a Species-Specific Virus Restriction Factor in Yeasts. PLoS Pathog. 2016, 12, e1005890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Rodríguez-Cousiño, N.; Esteban, R. Relationships and Evolution of Double-Stranded RNA Totiviruses of Yeasts Inferred from Analysis of L-A-2 and L-BC Variants in Wine Yeast Strain Populations. Appl. Environ. Microbiol. 2017, 83, e02991-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Rodríguez-Cousiño, N.; Gómez, P.; Esteban, R. L-A-Lus, a New Variant of the L-A Totivirus Found in Wine Yeasts with Klus Killer Toxin-Encoding Mlus Double-Stranded RNA: Possible Role of Killer Toxin-Encoding Satellite RNAs in the Evolution of Their Helper Viruses. Appl. Environ. Microbiol. 2013, 79, 4661–4674. [Google Scholar] [CrossRef] [Green Version]
  15. Konovalovas, A.; Serviené, E.; Serva, S. Genome Sequence of Saccharomyces cerevisiae Double-Stranded RNA Virus L-A-28. Genome Announc. 2016, 4, e00549-16. [Google Scholar] [CrossRef] [Green Version]
  16. Ramírez, M.; Velázquez, R.; López-Piñeiro, A.; Martínez, A. Genome Features of a New Double-Stranded RNA Helper Virus (LBCbarr) from Wine Torulaspora Delbrueckii Killer Strains. Int. J. Mol. Sci. 2021, 22, 13492. [Google Scholar] [CrossRef]
  17. Wickner, R.B. Double-Stranded and Single-Stranded RNA Viruses of Saccharomyces cerevisiae. Annu. Rev. Microbiol. 1992, 46, 347–375. [Google Scholar] [CrossRef]
  18. Aitmanaitė, L.; Konovalovas, A.; Medvedevas, P.; Servienė, E.; Serva, S. Specificity Determination in Saccharomyces cerevisiae Killer Virus Systems. Microorganisms 2021, 9, 236. [Google Scholar] [CrossRef]
  19. Sommer, S.S.; Wickner, R.B. Co-Curing of Plasmids Affecting Killer Double-Stranded RNAs of Saccharomyces cerevisiae: [HOK], [NEX], and the Abundance of L Are Related and Further Evidence That M1 Requires L. J. Bacteriol. 1982, 150, 545–551. [Google Scholar] [CrossRef] [Green Version]
  20. Esteban, R.; Fujimura, T.; Wickner, R.B. Internal and Terminal Cis-Acting Sites Are Necessary for In Vitro Replication of the L-A Double-Stranded RNA Virus of Yeast. EMBO J. 1989, 8, 947–954. [Google Scholar] [CrossRef]
  21. Fujimura, T.; Wickner, R.B. Replicase of L-A Virus-like Particles of Saccharomyces cerevisiae. In Vitro Conversion of Exogenous L-A and M1 Single-Stranded RNAs to Double-Stranded Form. J. Biol. Chem. 1988, 263, 454–460. [Google Scholar] [CrossRef]
  22. Fujimura, T.; Wickner, R.B. L-A Double-Stranded RNA Viruslike Particle Replication Cycle in Saccharomyces cerevisiae: Particle Maturation in Vitro and Effects of Mak10 and Pet18 Mutations. Mol. Cell. Biol. 1987, 7, 420–426. [Google Scholar] [CrossRef] [PubMed]
  23. Polevoda, B.; Sherman, F. NatC Nalpha-Terminal Acetyltransferase of Yeast Contains Three Subunits, Mak3p, Mak10p, and Mak31p. J. Biol. Chem. 2001, 276, 20154–20159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Meskauskas, A.; Petrov, A.N.; Dinman, J.D. Identification of Functionally Important Amino Acids of Ribosomal Protein L3 by Saturation Mutagenesis. Mol. Cell. Biol. 2005, 25, 10863–10874. [Google Scholar] [CrossRef] [Green Version]
  25. Ohtake, Y.; Wickner, R.B. Yeast Virus Propagation Depends Critically on Free 60S Ribosomal Subunit Concentration. Mol. Cell. Biol. 1995, 15, 2772–2781. [Google Scholar] [CrossRef] [Green Version]
  26. Meskauskas, A.; Baxter, J.L.; Carr, E.A.; Yasenchak, J.; Gallagher, J.E.G.; Baserga, S.J.; Dinman, J.D. Delayed RRNA Processing Results in Significant Ribosome Biogenesis and Functional Defects. Mol. Cell. Biol. 2003, 23, 1602–1613. [Google Scholar] [CrossRef] [Green Version]
  27. Meskauskas, A.; Harger, J.W.; Jacobs, K.L.M.; Dinman, J.D. Decreased Peptidyltransferase Activity Correlates with Increased Programmed-1 Ribosomal Frameshifting and Viral Maintenance Defects in the Yeast Saccharomyces cerevisiae. RNA 2003, 9, 982–992. [Google Scholar] [CrossRef] [Green Version]
  28. Dihanich, M.; van Tuinen, E.; Lambris, J.D.; Marshallsay, B. Accumulation of Viruslike Particles in a Yeast Mutant Lacking a Mitochondrial Pore Protein. Mol. Cell. Biol. 1989, 9, 1100–1108. [Google Scholar] [CrossRef]
  29. Oliver, S.G.; McCready, S.J.; Holm, C.; Sutherland, P.A.; McLaughlin, C.S.; Cox, B.S. Biochemical and Physiological Studies of the Yeast Virus-like Particle. J. Bacteriol. 1977, 130, 1303–1309. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, Y.J.; Wickner, R.B. MAK10, a Glucose-Repressible Gene Necessary for Replication of a DsRNA Virus of Saccharomyces cerevisiae, Has T Cell Receptor Alpha-Subunit Motifs. Genetics 1992, 132, 87–96. [Google Scholar] [CrossRef]
  31. Edwards, M.D.; Symbor-Nagrabska, A.; Dollard, L.; Gifford, D.K.; Fink, G.R. Interactions between Chromosomal and Nonchromosomal Elements Reveal Missing Heritability. Proc. Natl. Acad. Sci. USA 2014, 111, 7719–7722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Fujimura, T.; Esteban, R. Yeast Double-Stranded RNA Virus L-A Deliberately Synthesizes RNA Transcripts with 5′-Diphosphate. J. Biol. Chem. 2010, 285, 22911–22918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Widner, W.R.; Wickner, R.B. Evidence That the SKI Antiviral System of Saccharomyces cerevisiae Acts by Blocking Expression of Viral MRNA. Mol. Cell. Biol. 1993, 13, 4331–4341. [Google Scholar] [CrossRef] [PubMed]
  34. Solinger, J.A.; Pascolini, D.; Heyer, W.D. Active-Site Mutations in the Xrn1p Exoribonuclease of Saccharomyces cerevisiae Reveal a Specific Role in Meiosis. Mol. Cell. Biol. 1999, 19, 5930–5942. [Google Scholar] [CrossRef] [Green Version]
  35. Kim, J.; Kim, J. KEM1 Is Involved in Filamentous Growth of Saccharomyces cerevisiae. FEMS Microbiol. Lett. 2002, 216, 33–38. [Google Scholar] [CrossRef]
  36. Sheth, U.; Parker, R. Decapping and Decay of Messenger RNA Occur in Cytoplasmic Processing Bodies. Science 2003, 300, 805–808. [Google Scholar] [CrossRef] [Green Version]
  37. Askree, S.H.; Yehuda, T.; Smolikov, S.; Gurevich, R.; Hawk, J.; Coker, C.; Krauskopf, A.; Kupiec, M.; McEachern, M.J. A Genome-Wide Screen for Saccharomyces cerevisiae Deletion Mutants That Affect Telomere Length. Proc. Natl. Acad. Sci. USA 2004, 101, 8658–8663. [Google Scholar] [CrossRef] [Green Version]
  38. Sinturel, F.; Bréchemier-Baey, D.; Kiledjian, M.; Condon, C.; Bénard, L. Activation of 5′-3′ Exoribonuclease Xrn1 by Cofactor Dcs1 Is Essential for Mitochondrial Function in Yeast. Proc. Natl. Acad. Sci. USA 2012, 109, 8264–8269. [Google Scholar] [CrossRef] [Green Version]
  39. Delorme-Axford, E.; Abernathy, E.; Lennemann, N.J.; Bernard, A.; Ariosa, A.; Coyne, C.B.; Kirkegaard, K.; Klionsky, D.J. The Exoribonuclease Xrn1 Is a Post-Transcriptional Negative Regulator of Autophagy. Autophagy 2018, 14, 898–912. [Google Scholar] [CrossRef] [Green Version]
  40. Chlebowski, A.; Lubas, M.; Jensen, T.H.; Dziembowski, A. RNA Decay Machines: The Exosome. Biochim. Biophys. Acta 2013, 1829, 552–560. [Google Scholar] [CrossRef]
  41. Fujimura, T.; Esteban, R. Cap-Snatching Mechanism in Yeast L-A Double-Stranded RNA Virus. Proc. Natl. Acad. Sci. USA 2011, 108, 17667–17671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Georgopoulos, D.E.; Leibowitz, M.J. Nucleotide Phosphotransferase, Nucleotide Kinase and Inorganic Pyrophosphatase Activities of Killer Virions of Yeast. Yeast 1987, 3, 117–129. [Google Scholar] [CrossRef] [PubMed]
  43. Fujimura, T.; Esteban, R. Diphosphates at the 5′ End of the Positive Strand of Yeast L-A Double-Stranded RNA Virus as a Molecular Self-Identity Tag. Mol. Microbiol. 2016, 102, 71–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Liu, Y.X.; Dieckmann, C.L. Overproduction of Yeast Viruslike Particles by Strains Deficient in a Mitochondrial Nuclease. Mol. Cell. Biol. 1989, 9, 3323–3331. [Google Scholar] [CrossRef] [Green Version]
  45. Gao, J.; Chau, S.; Chowdhury, F.; Zhou, T.; Hossain, S.; McQuibban, G.A.; Meneghini, M.D. Meiotic Viral Attenuation through an Ancestral Apoptotic Pathway. Proc. Natl. Acad. Sci. USA 2019, 116, 16454–16462. [Google Scholar] [CrossRef] [Green Version]
  46. McBride, R.C.; Boucher, N.; Park, D.S.; Turner, P.E.; Townsend, J.P. Yeast Response to LA Virus Indicates Coadapted Global Gene Expression during Mycoviral Infection. FEMS Yeast Res. 2013, 13, 162–179. [Google Scholar] [CrossRef] [Green Version]
  47. Lukša, J.; Ravoitytė, B.; Konovalovas, A.; Aitmanaitė, L.; Butenko, A.; Yurchenko, V.; Serva, S.; Servienė, E. Different Metabolic Pathways Are Involved in Response of Saccharomyces cerevisiae to L-A and M Viruses. Toxins 2017, 9, 233. [Google Scholar] [CrossRef]
  48. Xu, S.; Yamamoto, N. Anti-Infective Nitazoxanide Disrupts Transcription of Ribosome Biogenesis-Related Genes in Yeast. Genes Genom. 2020, 42, 915–926. [Google Scholar] [CrossRef]
  49. Ravoitytė, B.; Lukša, J.; Yurchenko, V.; Serva, S.; Servienė, E. Saccharomyces paradoxus Transcriptional Alterations in Cells of Distinct Phenotype and Viral DsRNA Content. Microorganisms 2020, 8, 1902. [Google Scholar] [CrossRef]
  50. Pieczynska, M.D.; de Visser, J.A.G.M.; Korona, R. Incidence of Symbiotic DsRNA “killer” Viruses in Wild and Domesticated Yeast. FEMS Yeast Res. 2013, 13, 856–859. [Google Scholar] [CrossRef] [Green Version]
  51. Naumova, G.I.; Naumova, T.I. Comparative genetics of yeasts. XIII, Comparative study of Saccharomycetes-killers from different collections. Genetika 1973, 9, 140–145. [Google Scholar] [PubMed]
  52. Robinson, M.D.; Smyth, G.K. Small-Sample Estimation of Negative Binomial Dispersion, with Applications to SAGE Data. Biostatistics 2008, 9, 321–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Benjamini, Y.; Hochberg, Y. Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing. J. R. Stat. Soc. Ser. B 1995, 57, 289–300. [Google Scholar] [CrossRef]
  54. Balakrishnan, R.; Park, J.; Karra, K.; Hitz, B.C.; Binkley, G.; Hong, E.L.; Sullivan, J.; Micklem, G.; Cherry, J.M. YeastMine—An Integrated Data Warehouse for Saccharomyces cerevisiae Data as a Multipurpose Tool-Kit. Database 2012, 2012, bar062. [Google Scholar] [CrossRef]
  55. Boyle, E.I.; Weng, S.; Gollub, J.; Jin, H.; Botstein, D.; Cherry, J.M.; Sherlock, G. GO::TermFinder—Open Source Software for Accessing Gene Ontology Information and Finding Significantly Enriched Gene Ontology Terms Associated with a List of Genes. Bioinformatics 2004, 20, 3710–3715. [Google Scholar] [CrossRef] [Green Version]
  56. Cherry, J.M.; Hong, E.L.; Amundsen, C.; Balakrishnan, R.; Binkley, G.; Chan, E.T.; Christie, K.R.; Costanzo, M.C.; Dwight, S.S.; Engel, S.R.; et al. Saccharomyces Genome Database: The Genomics Resource of Budding Yeast. Nucleic Acids Res. 2012, 40, D700-5. [Google Scholar] [CrossRef] [Green Version]
  57. Monteiro, P.T.; Oliveira, J.; Pais, P.; Antunes, M.; Palma, M.; Cavalheiro, M.; Galocha, M.; Godinho, C.P.; Martins, L.C.; Bourbon, N.; et al. YEASTRACT+: A Portal for Cross-Species Comparative Genomics of Transcription Regulation in Yeasts. Nucleic Acids Res. 2020, 48, D642–D649. [Google Scholar] [CrossRef]
  58. Fernández-Cid, A.; Riera, A.; Herrero, P.; Moreno, F. Glucose Levels Regulate the Nucleo-Mitochondrial Distribution of Mig2. Mitochondrion 2012, 12, 370–380. [Google Scholar] [CrossRef]
  59. Zhao, H.; Eide, D. The Yeast ZRT1 Gene Encodes the Zinc Transporter Protein of a High-Affinity Uptake System Induced by Zinc Limitation. Proc. Natl. Acad. Sci. USA 1996, 93, 2454–2458. [Google Scholar] [CrossRef] [Green Version]
  60. Brzywczy, J.; Paszewski, A. Role of O-Acetylhomoserine Sulfhydrylase in Sulfur Amino Acid Synthesis in Various Yeasts. Yeast 1993, 9, 1335–1342. [Google Scholar] [CrossRef]
  61. Gier, S.; Simon, M.; Gasparoni, G.; Khalifa, S.; Schulz, M.H.; Schmitt, M.J.; Breinig, F. Yeast Viral Killer Toxin K1 Induces Specific Host Cell Adaptions via Intrinsic Selection Pressure. Appl. Environ. Microbiol. 2020, 86, e02446-19. [Google Scholar] [CrossRef] [PubMed]
  62. Becker, B.; Schmitt, M.J. Yeast Killer Toxin K28: Biology and Unique Strategy of Host Cell Intoxication and Killing. Toxins 2017, 9, 333. [Google Scholar] [CrossRef] [PubMed]
  63. Tipper, D.J.; Schmitt, M.J. Yeast DsRNA Viruses: Replication and Killer Phenotypes. Mol. Microbiol. 1991, 5, 2331–2338. [Google Scholar] [CrossRef] [PubMed]
  64. Day, M. Yeast Petites and Small Colony Variants: For Everything There Is a Season. Adv. Appl. Microbiol. 2013, 85, 1–41. [Google Scholar] [CrossRef]
  65. Evans, I.H.; Diala, E.S.; Earl, A.; Wilkie, D. Mitochondrial Control of Cell Surface Characteristics in Saccharomyces cerevisiae. Biochim. Biophys. Acta 1980, 602, 201–206. [Google Scholar] [CrossRef]
  66. Traven, A.; Wong, J.M.; Xu, D.; Sopta, M.; Ingles, C.J. Interorganellar Communication. Altered Nuclear Gene Expression Profiles in a Yeast Mitochondrial DNA Mutant. J. Biol. Chem. 2001, 276, 4020–4027. [Google Scholar] [CrossRef] [Green Version]
  67. Kennedy, B.K.; Austriaco, N.R.J.; Zhang, J.; Guarente, L. Mutation in the Silencing Gene SIR4 Can Delay Aging in S. cerevisiae. Cell 1995, 80, 485–496. [Google Scholar] [CrossRef] [Green Version]
  68. Kispal, G.; Sipos, K.; Lange, H.; Fekete, Z.; Bedekovics, T.; Janáky, T.; Bassler, J.; Aguilar Netz, D.J.; Balk, J.; Rotte, C.; et al. Biogenesis of Cytosolic Ribosomes Requires the Essential Iron-Sulphur Protein Rli1p and Mitochondria. EMBO J. 2005, 24, 589–598. [Google Scholar] [CrossRef] [Green Version]
  69. Wickner, R.B. Chromosomal and Nonchromosomal Mutations Affecting the “Killer Character” of Saccharomyces cerevisiae. Genetics 1974, 76, 423–432. [Google Scholar] [CrossRef]
  70. Unchwaniwala, N.; Zhan, H.; Pennington, J.; Horswill, M.; den Boon, J.A.; Ahlquist, P. Subdomain Cryo-EM Structure of Nodaviral Replication Protein A Crown Complex Provides Mechanistic Insights into RNA Genome Replication. Proc. Natl. Acad. Sci. USA 2020, 117, 18680–18691. [Google Scholar] [CrossRef]
  71. Lau, S.K.P.; Lo, G.C.S.; Chow, F.W.N.; Fan, R.Y.Y.; Cai, J.J.; Yuen, K.-Y.; Woo, P.C.Y. Novel Partitivirus Enhances Virulence of and Causes Aberrant Gene Expression in Talaromyces marneffei. mBio 2018, 9, e00947-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Drinnenberg, I.A.; Fink, G.R.; Bartel, D.P. Compatibility with Killer Explains the Rise of RNAi-Deficient Fungi. Science 2011, 333, 1592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Clancey, S.A.; Ruchti, F.; LeibundGut-Landmann, S.; Heitman, J.; Ianiri, G. A Novel Mycovirus Evokes Transcriptional Rewiring in the Fungus Malassezia and Stimulates Interferon-β Production in Macrophages. mBio 2020, 11, e01534-20. [Google Scholar] [CrossRef]
  74. Park, M.; Cho, Y.-J.; Kim, D.; Yang, C.-S.; Lee, S.M.; Dawson, T.L.; Nakamizo, S.; Kabashima, K.; Lee, Y.W.; Jung, W.H. A Novel Virus Alters Gene Expression and Vacuolar Morphology in Malassezia Cells and Induces a TLR3-Mediated Inflammatory Immune Response. mBio 2020, 11, e01521-20. [Google Scholar] [CrossRef] [PubMed]
  75. Lee Marzano, S.-Y.; Neupane, A.; Domier, L. Transcriptional and Small RNA Responses of the White Mold Fungus Sclerotinia sclerotiorum to Infection by a Virulence-Attenuating Hypovirus. Viruses 2018, 10, 713. [Google Scholar] [CrossRef] [Green Version]
  76. Adaui, V.; Lye, L.-F.; Akopyants, N.S.; Zimic, M.; Llanos-Cuentas, A.; Garcia, L.; Maes, I.; De Doncker, S.; Dobson, D.E.; Arevalo, J.; et al. Association of the Endobiont Double-Stranded RNA Virus LRV1 With Treatment Failure for Human Leishmaniasis Caused by Leishmania braziliensis in Peru and Bolivia. J. Infect. Dis. 2016, 213, 112–121. [Google Scholar] [CrossRef] [Green Version]
  77. Grybchuk, D.; Akopyants, N.S.; Kostygov, A.Y.; Konovalovas, A.; Lye, L.-F.; Dobson, D.E.; Zangger, H.; Fasel, N.; Butenko, A.; Frolov, A.O.; et al. Viral Discovery and Diversity in Trypanosomatid Protozoa with a Focus on Relatives of the Human Parasite Leishmania. Proc. Natl. Acad. Sci. USA 2018, 115, E506–E515. [Google Scholar] [CrossRef] [Green Version]
  78. Kotta-Loizou, I. Mycoviruses and Their Role in Fungal Pathogenesis. Curr. Opin. Microbiol. 2021, 63, 10–18. [Google Scholar] [CrossRef]
  79. Święciło, A. Cross-Stress Resistance in Saccharomyces cerevisiae Yeast--New Insight into an Old Phenomenon. Cell Stress Chaperones 2016, 21, 187–200. [Google Scholar] [CrossRef] [Green Version]
  80. Sánchez-Gaya, V.; Casaní-Galdón, S.; Ugidos, M.; Kuang, Z.; Mellor, J.; Conesa, A.; Tarazona, S. Elucidating the Role of Chromatin State and Transcription Factors on the Regulation of the Yeast Metabolic Cycle: A Multi-Omic Integrative Approach. Front. Genet. 2018, 9, 578. [Google Scholar] [CrossRef] [Green Version]
  81. Karunanithi, S.; Cullen, P.J. The Filamentous Growth MAPK Pathway Responds to Glucose Starvation Through the Mig1/2 Transcriptional Repressors in Saccharomyces cerevisiae. Genetics 2012, 192, 869–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Angus-Hill, M.L.; Schlichter, A.; Roberts, D.; Erdjument-Bromage, H.; Tempst, P.; Cairns, B.R. A Rsc3/Rsc30 Zinc Cluster Dimer Reveals Novel Roles for the Chromatin Remodeler RSC in Gene Expression and Cell Cycle Control. Mol. Cell 2001, 7, 741–751. [Google Scholar] [CrossRef]
  83. Hill, M.S.; Vande Zande, P.; Wittkopp, P.J. Molecular and Evolutionary Processes Generating Variation in Gene Expression. Nat. Rev. Genet. 2021, 22, 203–215. [Google Scholar] [CrossRef] [PubMed]
  84. Hahn, S.; Young, E.T. Transcriptional Regulation in Saccharomyces cerevisiae: Transcription Factor Regulation and Function, Mechanisms of Initiation, and Roles of Activators and Coactivators. Genetics 2011, 189, 705–736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Hughes, T.R.; de Boer, C.G. Mapping Yeast Transcriptional Networks. Genetics 2013, 195, 9–36. [Google Scholar] [CrossRef] [Green Version]
  86. Reimand, J.; Vaquerizas, J.M.; Todd, A.E.; Vilo, J.; Luscombe, N.M. Comprehensive Reanalysis of Transcription Factor Knockout Expression Data in Saccharomyces cerevisiae Reveals Many New Targets. Nucleic Acids Res. 2010, 38, 4768–4777. [Google Scholar] [CrossRef] [Green Version]
  87. Marion, R.M.; Regev, A.; Segal, E.; Barash, Y.; Koller, D.; Friedman, N.; O’Shea, E.K. Sfp1 Is a Stress- and Nutrient-Sensitive Regulator of Ribosomal Protein Gene Expression. Proc. Natl. Acad. Sci. USA 2004, 101, 14315–14322. [Google Scholar] [CrossRef] [Green Version]
  88. Jorgensen, P.; Rupes, I.; Sharom, J.R.; Schneper, L.; Broach, J.R.; Tyers, M. A Dynamic Transcriptional Network Communicates Growth Potential to Ribosome Synthesis and Critical Cell Size. Genes Dev. 2004, 18, 2491–2505. [Google Scholar] [CrossRef] [Green Version]
  89. Gertz, J.; Riles, L.; Turnbaugh, P.; Ho, S.-W.; Cohen, B.A. Discovery, Validation, and Genetic Dissection of Transcription Factor Binding Sites by Comparative and Functional Genomics. Genome Res. 2005, 15, 1145–1152. [Google Scholar] [CrossRef] [Green Version]
  90. Cipollina, C.; van den Brink, J.; Daran-Lapujade, P.; Pronk, J.T.; Porro, D.; de Winde, J.H. Saccharomyces cerevisiaeSFP1: At the Crossroads of Central Metabolism and Ribosome Biogenesis. Microbiology 2008, 154, 1686–1699. [Google Scholar] [CrossRef] [Green Version]
  91. Venters, B.J.; Wachi, S.; Mavrich, T.N.; Andersen, B.E.; Jena, P.; Sinnamon, A.J.; Jain, P.; Rolleri, N.S.; Jiang, C.; Hemeryck-Walsh, C.; et al. A Comprehensive Genomic Binding Map of Gene and Chromatin Regulatory Proteins in Saccharomyces. Mol. Cell 2011, 41, 480–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Vyas, V.K.; Berkey, C.D.; Miyao, T.; Carlson, M. Repressors Nrg1 and Nrg2 Regulate a Set of Stress-Responsive Genes in Saccharomyces cerevisiae. Eukaryot. Cell 2005, 4, 1882–1891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. MacPherson, S.; Larochelle, M.; Turcotte, B. A Fungal Family of Transcriptional Regulators: The Zinc Cluster Proteins. Microbiol. Mol. Biol. Rev. 2006, 70, 583–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Friden, P.; Schimmel, P. LEU3 of Saccharomyces cerevisiae Activates Multiple Genes for Branched-Chain Amino Acid Biosynthesis by Binding to a Common Decanucleotide Core Sequence. Mol. Cell. Biol. 1988, 8, 2690–2697. [Google Scholar] [CrossRef] [PubMed]
  95. dos Santos, S.C.; Tenreiro, S.; Palma, M.; Becker, J.; Sá-Correia, I. Transcriptomic Profiling of the Saccharomyces cerevisiae Response to Quinine Reveals a Glucose Limitation Response Attributable to Drug-Induced Inhibition of Glucose Uptake. Antimicrob. Agents Chemother. 2009, 53, 5213–5223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Gagiano, M.; Bester, M.; van Dyk, D.; Franken, J.; Bauer, F.F.; Pretorius, I.S. Mss11p Is a Transcription Factor Regulating Pseudohyphal Differentiation, Invasive Growth and Starch Metabolism in Saccharomyces cerevisiae in Response to Nutrient Availability. Mol. Microbiol. 2003, 47, 119–134. [Google Scholar] [CrossRef] [PubMed]
  97. Ross, C.D.; McCarty, B.R.; Hamilton, M.; Ben-Hur, A.; Ross, E.D. A Promiscuous Prion: Efficient Induction of [URE3] Prion Formation by Heterologous Prion Domains. Genetics 2009, 183, 929–940. [Google Scholar] [CrossRef] [Green Version]
  98. Thepnok, P.; Ratanakhanokchai, K.; Soontorngun, N. The Novel Zinc Cluster Regulator Tog1 Plays Important Roles in Oleate Utilization and Oxidative Stress Response in Saccharomyces cerevisiae. Biochem. Biophys. Res. Commun. 2014, 450, 1276–1282. [Google Scholar] [CrossRef]
  99. Uemura, H.; Jigami, Y. Mutations in GCR1, a Transcriptional Activator of Saccharomyces cerevisiae Glycolytic Genes, Function as Suppressors of Gcr2 Mutations. Genetics 1995, 139, 511–521. [Google Scholar] [CrossRef]
  100. Kratzer, S.; Schüller, H.J. Transcriptional Control of the Yeast Acetyl-CoA Synthetase Gene, ACS1, by the Positive Regulators CAT8 and ADR1 and the Pleiotropic Repressor UME6. Mol. Microbiol. 1997, 26, 631–641. [Google Scholar] [CrossRef]
  101. Gaspar, M.L.; Chang, Y.-F.; Jesch, S.A.; Aregullin, M.; Henry, S.A. Interaction between Repressor Opi1p and ER Membrane Protein Scs2p Facilitates Transit of Phosphatidic Acid from the ER to Mitochondria and Is Essential for INO1 Gene Expression in the Presence of Choline. J. Biol. Chem. 2017, 292, 18713–18728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Buechel, E.R.; Pinkett, H.W. Transcription Factors and ABC Transporters: From Pleiotropic Drug Resistance to Cellular Signaling in Yeast. FEBS Lett. 2020, 594, 3943–3964. [Google Scholar] [CrossRef] [PubMed]
  103. Shu, W.-J.; Chen, R.; Yin, Z.-H.; Li, F.; Zhang, H.; Du, H.-N. Rph1 Coordinates Transcription of Ribosomal Protein Genes and Ribosomal RNAs to Control Cell Growth under Nutrient Stress Conditions. Nucleic Acids Res. 2020, 48, 8360–8373. [Google Scholar] [CrossRef] [PubMed]
  104. Pietrocola, F.; Galluzzi, L.; Bravo-San Pedro, J.M.; Madeo, F.; Kroemer, G. Acetyl Coenzyme A: A Central Metabolite and Second Messenger. Cell Metab. 2015, 21, 805–821. [Google Scholar] [CrossRef] [Green Version]
  105. Galdieri, L.; Zhang, T.; Rogerson, D.; Lleshi, R.; Vancura, A. Protein Acetylation and Acetyl Coenzyme a Metabolism in Budding Yeast. Eukaryot. Cell 2014, 13, 1472–1483. [Google Scholar] [CrossRef] [Green Version]
  106. Cartwright, J.L.; Gasmi, L.; Spiller, D.G.; McLennan, A.G. The Saccharomyces cerevisiae PCD1 Gene Encodes a Peroxisomal Nudix Hydrolase Active toward Coenzyme A and Its Derivatives. J. Biol. Chem. 2000, 275, 32925–32930. [Google Scholar] [CrossRef] [Green Version]
  107. Kayikci, Ö.; Nielsen, J. Glucose Repression in Saccharomyces cerevisiae. FEMS Yeast Res. 2015, 15, fov068. [Google Scholar] [CrossRef] [Green Version]
  108. Zhang, M.; Galdieri, L.; Vancura, A. The Yeast AMPK Homolog SNF1 Regulates Acetyl Coenzyme A Homeostasis and Histone Acetylation. Mol. Cell. Biol. 2013, 33, 4701–4717. [Google Scholar] [CrossRef] [Green Version]
  109. Grunwald, S.; Hopf, L.V.M.; Bock-Bierbaum, T.; Lally, C.C.M.; Spahn, C.M.T.; Daumke, O. Divergent Architecture of the Heterotrimeric NatC Complex Explains N-Terminal Acetylation of Cognate Substrates. Nat. Commun. 2020, 11, 5506. [Google Scholar] [CrossRef]
  110. Tercero, J.C.; Dinman, J.D.; Wickner, R.B. Yeast MAK3 N-Acetyltransferase Recognizes the N-Terminal Four Amino Acids of the Major Coat Protein (Gag) of the L-A Double-Stranded RNA Virus. J. Bacteriol. 1993, 175, 3192–3194. [Google Scholar] [CrossRef] [Green Version]
  111. Drazic, A.; Varland, S. Human NAA30 Can Rescue Yeast Mak3∆ Mutant Growth Phenotypes. Biosci. Rep. 2021, 41, BSR20202828. [Google Scholar] [CrossRef] [PubMed]
  112. Aksnes, H.; Ree, R.; Arnesen, T. Co-Translational, Post-Translational, and Non-Catalytic Roles of N-Terminal Acetyltransferases. Mol. Cell 2019, 73, 1097–1114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kalvik, T.V.; Arnesen, T. Protein N-Terminal Acetyltransferases in Cancer. Oncogene 2013, 32, 269–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Le Crom, S.; Devaux, F.; Marc, P.; Zhang, X.; Moye-Rowley, W.S.; Jacq, C. New Insights into the Pleiotropic Drug Resistance Network from Genome-Wide Characterization of the YRR1 Transcription Factor Regulation System. Mol. Cell. Biol. 2002, 22, 2642–2649. [Google Scholar] [CrossRef] [Green Version]
  115. Jiang, Y.; Vasconcelles, M.J.; Wretzel, S.; Light, A.; Martin, C.E.; Goldberg, M.A. MGA2 Is Involved in the Low-Oxygen Response Element-Dependent Hypoxic Induction of Genes in Saccharomyces cerevisiae. Mol. Cell. Biol. 2001, 21, 6161–6169. [Google Scholar] [CrossRef] [Green Version]
  116. Romero, A.M.; Jordá, T.; Rozès, N.; Martínez-Pastor, M.T.; Puig, S. Regulation of Yeast Fatty Acid Desaturase in Response to Iron Deficiency. Biochim. Biophys. Acta. Mol. Cell Biol. Lipids 2018, 1863, 657–668. [Google Scholar] [CrossRef] [Green Version]
  117. Bourot, S.; Karst, F. Isolation and Characterization of the Saccharomyces cerevisiae SUT1 Gene Involved in Sterol Uptake. Gene 1995, 165, 97–102. [Google Scholar] [CrossRef]
  118. Regnacq, M.; Alimardani, P.; El Moudni, B.; Berges, T. SUT1p Interaction with Cyc8p(Ssn6p) Relieves Hypoxic Genes from Cyc8p-Tup1p Repression in Saccharomyces cerevisiae. Mol. Microbiol. 2001, 40, 1085–1096. [Google Scholar] [CrossRef] [Green Version]
  119. Foster, H.A.; Cui, M.; Naveenathayalan, A.; Unden, H.; Schwanbeck, R.; Hofken, T. The Zinc Cluster Protein Sut1 Contributes to Filamentation in Saccharomyces cerevisiae. Eukaryot. Cell 2013, 12, 244–253. [Google Scholar] [CrossRef] [Green Version]
  120. Blanda, C.; Hofken, T. Regulation of Mating in the Budding Yeast Saccharomyces cerevisiae by the Zinc Cluster Proteins Sut1 and Sut2. Biochem. Biophys. Res. Commun. 2013, 438, 66–70. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Volcano Plots for differential gene expression in S. cerevisiae M437[L] and two S. paradoxus AML-15-66 strains—Spo[L] and Pet[L]. Scattered points represent differently expressed genes (DEGs): red, upregulated; blue, downregulated. The x-axis represents the fold-change of the DEGs, whereas the y-axis indicates the log odds—the probability that the differential expression of a gene has statistical significance. DEGs in (A) M437[L], (B) Spo[L], and (C) Pet[L] cells are represented. M437[L], Spo[L], Pet[L]: DEGs in solely L-A-maintaining cells with reference to dsRNA-free cells of the same lineage.
Figure 1. Volcano Plots for differential gene expression in S. cerevisiae M437[L] and two S. paradoxus AML-15-66 strains—Spo[L] and Pet[L]. Scattered points represent differently expressed genes (DEGs): red, upregulated; blue, downregulated. The x-axis represents the fold-change of the DEGs, whereas the y-axis indicates the log odds—the probability that the differential expression of a gene has statistical significance. DEGs in (A) M437[L], (B) Spo[L], and (C) Pet[L] cells are represented. M437[L], Spo[L], Pet[L]: DEGs in solely L-A-maintaining cells with reference to dsRNA-free cells of the same lineage.
Jof 08 00381 g001
Figure 2. Venn diagrams representing the numbers of differently expressed genes (DEGs) in S. cerevisiae M437[L], and S. paradoxus AML-15-66 Spo[L] and Pet[L] cells. M437[L], Spo[L], and Pet[L]: differently expressed genes in solely L-A maintaining cells with reference to dsRNA-free cells of the same lineage.
Figure 2. Venn diagrams representing the numbers of differently expressed genes (DEGs) in S. cerevisiae M437[L], and S. paradoxus AML-15-66 Spo[L] and Pet[L] cells. M437[L], Spo[L], and Pet[L]: differently expressed genes in solely L-A maintaining cells with reference to dsRNA-free cells of the same lineage.
Jof 08 00381 g002
Figure 3. Selected statistically significant enriched gene ontology terms associated with the biological processes of altered transcription genes in S. cerevisiae M437[L] cells. The green line represents the number of DEGs, whereas the bars indicate the fold-enrichment (F.E.) values. Upregulated genes, red; downregulated genes, blue. F.E. values were calculated by dividing the frequency of the specific gene cluster by the total frequency for each GO term according to the data presented in Table S1.
Figure 3. Selected statistically significant enriched gene ontology terms associated with the biological processes of altered transcription genes in S. cerevisiae M437[L] cells. The green line represents the number of DEGs, whereas the bars indicate the fold-enrichment (F.E.) values. Upregulated genes, red; downregulated genes, blue. F.E. values were calculated by dividing the frequency of the specific gene cluster by the total frequency for each GO term according to the data presented in Table S1.
Jof 08 00381 g003
Figure 4. Selected statistically significant enriched gene ontology terms associated with the biological processes of altered transcription genes in (A) S. paradoxus AML-15-66 Spo[L] and (B) Pet[L] cells. The green line represents the number of DEGs, whereas the bars indicate fold enrichment (F.E.) values. Upregulated genes, red; downregulated genes, blue. F.E. values were calculated by dividing the frequency of a specific gene cluster by the total frequency for each GO term according to the data presented in Table S1.
Figure 4. Selected statistically significant enriched gene ontology terms associated with the biological processes of altered transcription genes in (A) S. paradoxus AML-15-66 Spo[L] and (B) Pet[L] cells. The green line represents the number of DEGs, whereas the bars indicate fold enrichment (F.E.) values. Upregulated genes, red; downregulated genes, blue. F.E. values were calculated by dividing the frequency of a specific gene cluster by the total frequency for each GO term according to the data presented in Table S1.
Jof 08 00381 g004
Figure 5. Associations between differently expressed genes (DEGs) and their documented regulators in S. cerevisiae M437[L], and S. paradoxus AML-15-66 Spo[L] and Pet[L] cells. Upregulated genes were only searched for transcriptional activators, while downregulated genes were only searched for transcriptional repressors using the tools available in the YEASTRACT database. TF: transcription factor; M437[L], Spo[L], Pet[L]: DEGs in solely L-A maintaining cells with reference to dsRNA-free cells of the same lineage. The impact of each TF was predicted based on the percentage of targets in each dataset only considering documented expression evidence. The percentage of genes regulated by each TF in each dataset of up- or downregulated genes is presented in the table. Only selected highest-confidence-level regulatory associations are presented. The higher green color intensity indicates the higher percentage of regulated genes. The full list is available in Table S2.
Figure 5. Associations between differently expressed genes (DEGs) and their documented regulators in S. cerevisiae M437[L], and S. paradoxus AML-15-66 Spo[L] and Pet[L] cells. Upregulated genes were only searched for transcriptional activators, while downregulated genes were only searched for transcriptional repressors using the tools available in the YEASTRACT database. TF: transcription factor; M437[L], Spo[L], Pet[L]: DEGs in solely L-A maintaining cells with reference to dsRNA-free cells of the same lineage. The impact of each TF was predicted based on the percentage of targets in each dataset only considering documented expression evidence. The percentage of genes regulated by each TF in each dataset of up- or downregulated genes is presented in the table. Only selected highest-confidence-level regulatory associations are presented. The higher green color intensity indicates the higher percentage of regulated genes. The full list is available in Table S2.
Jof 08 00381 g005
Figure 6. Relative L-A dsRNA content in S. cerevisiae cells upon the deletion or overexpression of single genes. Strains having increased or decreased amounts of L-A dsRNA are depicted; wt indicates levels similar to the parental BY4741 strain. The dsRNA levels of the L-A virus were normalized by 18S rRNA and compared to the wt control. L-A dsRNA levels in transcription factor (TF)-deletion strains (A), in strains overexpressing TF (B), and in strains lacking genes linked to the acetyl-CoA levels (C).
Figure 6. Relative L-A dsRNA content in S. cerevisiae cells upon the deletion or overexpression of single genes. Strains having increased or decreased amounts of L-A dsRNA are depicted; wt indicates levels similar to the parental BY4741 strain. The dsRNA levels of the L-A virus were normalized by 18S rRNA and compared to the wt control. L-A dsRNA levels in transcription factor (TF)-deletion strains (A), in strains overexpressing TF (B), and in strains lacking genes linked to the acetyl-CoA levels (C).
Jof 08 00381 g006
Figure 7. Schematic representation of links between transcription factors (TFs), differently expressed genes (DEGs), and levels of L-A virus. (A) TF can either activate (orange and green TFs) or inhibit (blue and yellow TFs) the transcription of genes. DEGs can either favor or repress the propagation of L-A virus, thereby resulting in alterations of the L-A dsRNA levels. Consequently, the deletion (ΔTF) or overexpression (TFOE) of certain TFs can contribute to changes in gene expression. The elimination or overexpression of certain TFs can change—increase (+) or decrease (−)—the intracellular L-A dsRNA level by changing the balance of transcripts that are important for L-A virus maintenance. All TFs act simultaneously to impact the total amount of L-A virus. (B) Relationship between TFs and known cellular factors affecting L-A virus replication. TFs regulate the transcription of genes involved in nutrient metabolism, which is directly linked to the cellular metabolite levels and ribosome biogenesis. Acetyl-CoA is a substrate for the NatC acetyltransferase complex, composed of Mak3, Mak10, and Mak31 subunits, which is necessary for co-translational Gag protein acetylation. The translation of viral proteins by ribosome and Gag acetylation is essential for L-A virus assembly and propagation.
Figure 7. Schematic representation of links between transcription factors (TFs), differently expressed genes (DEGs), and levels of L-A virus. (A) TF can either activate (orange and green TFs) or inhibit (blue and yellow TFs) the transcription of genes. DEGs can either favor or repress the propagation of L-A virus, thereby resulting in alterations of the L-A dsRNA levels. Consequently, the deletion (ΔTF) or overexpression (TFOE) of certain TFs can contribute to changes in gene expression. The elimination or overexpression of certain TFs can change—increase (+) or decrease (−)—the intracellular L-A dsRNA level by changing the balance of transcripts that are important for L-A virus maintenance. All TFs act simultaneously to impact the total amount of L-A virus. (B) Relationship between TFs and known cellular factors affecting L-A virus replication. TFs regulate the transcription of genes involved in nutrient metabolism, which is directly linked to the cellular metabolite levels and ribosome biogenesis. Acetyl-CoA is a substrate for the NatC acetyltransferase complex, composed of Mak3, Mak10, and Mak31 subunits, which is necessary for co-translational Gag protein acetylation. The translation of viral proteins by ribosome and Gag acetylation is essential for L-A virus assembly and propagation.
Jof 08 00381 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ravoitytė, B.; Lukša, J.; Wellinger, R.E.; Serva, S.; Servienė, E. Adaptive Response of Saccharomyces Hosts to Totiviridae L-A dsRNA Viruses Is Achieved through Intrinsically Balanced Action of Targeted Transcription Factors. J. Fungi 2022, 8, 381. https://doi.org/10.3390/jof8040381

AMA Style

Ravoitytė B, Lukša J, Wellinger RE, Serva S, Servienė E. Adaptive Response of Saccharomyces Hosts to Totiviridae L-A dsRNA Viruses Is Achieved through Intrinsically Balanced Action of Targeted Transcription Factors. Journal of Fungi. 2022; 8(4):381. https://doi.org/10.3390/jof8040381

Chicago/Turabian Style

Ravoitytė, Bazilė, Juliana Lukša, Ralf Erik Wellinger, Saulius Serva, and Elena Servienė. 2022. "Adaptive Response of Saccharomyces Hosts to Totiviridae L-A dsRNA Viruses Is Achieved through Intrinsically Balanced Action of Targeted Transcription Factors" Journal of Fungi 8, no. 4: 381. https://doi.org/10.3390/jof8040381

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