Next Article in Journal
β-Glucan Induces Training Immunity to Promote Antiviral Activity by Activating TBK1
Next Article in Special Issue
Temperature Effects on the Cryphonectria hypovirus 1 Accumulation and Recovery within Its Fungal Host, the Chestnut Blight Pathogen Cryphonectria parasitica
Previous Article in Journal
Poly(I:C) Induces Distinct Liver Cell Type-Specific Responses in Hepatitis B Virus-Transgenic Mice In Vitro, but Fails to Induce These Signals In Vivo
Previous Article in Special Issue
Virus Infection Impairs Fungal Response to Stress: Effect of Salt
 
 
Correction published on 19 April 2024, see Viruses 2024, 16(4), 632.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fungal Viruses Unveiled: A Comprehensive Review of Mycoviruses

Forestry & Agricultural Biotechnology Institute (FABI), Department of Biochemistry, Genetics & Microbiology, University of Pretoria, Pretoria 0002, South Africa
*
Author to whom correspondence should be addressed.
Viruses 2023, 15(5), 1202; https://doi.org/10.3390/v15051202
Submission received: 14 April 2023 / Revised: 7 May 2023 / Accepted: 17 May 2023 / Published: 19 May 2023 / Corrected: 19 April 2024
(This article belongs to the Collection Mycoviruses)

Abstract

:
Mycoviruses (viruses of fungi) are ubiquitous throughout the fungal kingdom and are currently classified into 23 viral families and the genus botybirnavirus by the International Committee on the Taxonomy of Viruses (ICTV). The primary focus of mycoviral research has been on mycoviruses that infect plant pathogenic fungi, due to the ability of some to reduce the virulence of their host and thus act as potential biocontrol against these fungi. However, mycoviruses lack extracellular transmission mechanisms and rely on intercellular transmission through the hyphal anastomosis, which impedes successful transmission between different fungal strains. This review provides a comprehensive overview of mycoviruses, including their origins, host range, taxonomic classification into families, effects on their fungal counterparts, and the techniques employed in their discovery. The application of mycoviruses as biocontrol agents of plant pathogenic fungi is also discussed.

1. Introduction

Mycoviruses (viruses of fungi) are ubiquitous throughout the fungal kingdom [1]. They are known to associate with most of the major fungal taxonomic groups, including Ascomycota, Basidiomycota, Chytridiomycota, Zygomycota, and Neocallimastigomycota [2,3]. The International Committee on the Taxonomy of Viruses (ICTV) currently classifies mycoviruses as 23 families and 1 unclassified genus, and this is based on the type of genome [4,5,6,7,8]. The majority of mycoviruses have double-stranded RNA (dsRNA) or positive sense single-stranded RNA (+ssRNA) genomes [5]. Mycoviruses with dsRNA genomes are classified into the viral families Chrysoviridae, Amalgaviridae Megabirnaviridae, Quadriviridae, Partitiviridae, Polymycoviridae, Reoviridae, Totiviridae, and the genus Botybirnavirus (unclassified) [4,6]. Those with +ssRNA genomes are grouped into the families Endornaviridae, Alphaflexiviridae, Barnaviridae, Deltaflexiviridae, Gammaflexiviridae, Hypoviridae, Narnaviridae, Mitoviridae, Hadakaviridae, Yadokariviridae, and the reverse transcribing (RT) families Metaviridae and Pseudoviridae [4,6,7,8]. Mycoviruses with negative-sense single-stranded RNA (−ssRNA) genomes have also been discovered [9,10,11], and belong to the family Mymonaviridae [4,6]. Recently a number of ssDNA mycoviruses have also been found; however, only two belong to a recognized mycoviral family, namely Sclerotinia sclerotiorum hypovirulence-associated DNA virus 1 and Fusarium graminearum gemytripvirus 1 [12]. Both belong to the family Genomoviridae [13,14].
Mycoviruses that elicit hypovirulence have been identified in both human and plant pathogenic fungi. These viruses frequently cause several adverse effects in their fungal hosts, such as decreased virulence, irregular growth, abnormal pigmentation, and defects in sexual development [5,15,16]. The recent discovery of hypovirulence inducing mycoviruses in human pathogenic fungi presents an opportunity for the development of therapeutic interventions against fungal infections in humans. The majority of mycoviral research, however, has been concerned with hypovirulence-inducing mycoviruses of plant pathogenic fungi [12,17,18]. These mycoviruses have the potential to be used as biocontrol agents against their fungal hosts, thereby reducing the losses in agriculture and forestry due to fungal infections [19,20]. Despite their potential, the use of mycoviruses as biocontrol agents has a number of challenges. Mycoviruses lack the extracellular transmission mechanisms of plant and animal viruses [3]. Instead, viral transmission occurs intercellularly, through hyphal anastomosis (fusion of fungal hyphae), cell division, and sporulation [3]. Hyphal anastomosis impedes the successful transmission of mycoviruses between different fungal strains since they need to be vegetatively compatible with hyphal fusion [21]. Nevertheless, research is ongoing to find solutions to this problem [22,23], and mycovirus-based biocontrol has already been used successfully [19].
Advances in high throughput sequencing have led to a surge in mycovirus discoveries and new insights into their origins, diversity, and impact on fungal hosts. We review the latest developments in the field of mycovirology and evaluate the potential of mycoviruses as biocontrol agents of plant pathogenic fungi. We also summarize some of the methods which have been employed to help aid the discovery process.

2. The History and Origins of Mycoviruses

In 1948, a disease-causing dieback of the commercially produced mushroom, Agaricus bisporus, was reported and named the ‘La France’ disease [24]. It was not until nearly a decade later that mycoviruses were identified as the causal agents of this economically important disease [25]. This quickly led to the development of a new field of study: mycovirology. Not long after, mycoviruses were also identified in the ascomycete Penicillium stoloniferum, and were determined to cause interferon stimulation in mammals [26,27]. It was not until the 1970s, however, that a significant breakthrough was made in the field of mycovirology. A mycovirus of the chestnut blight pathogen, Cryphonectria parasitica, was observed to reduce the virulence of its host and had potential as a biocontrol agent against this plant pathogenic fungus [28]. This led to increased interest in the discovery of mycoviruses in plant pathogenic fungi, as they could serve as prospective biocontrol agents of such fungi. Another development during this period was the discovery of mycoviruses which induced the ‘killer yeast’ phenotype in Saccharomyces cerevisiae, a fungus that is routinely employed in winemaking, brewing, and baking [29]. These mycoviruses confer a competitive advantage to their fungal hosts by producing extracellular toxins which eliminate competing strains [30]. While these toxins were initially associated with fermentation failure, their potential use as a means of eliminating undesirable strains was soon realized [30].
Recent advances in high throughput technologies, specifically RNA sequencing, have facilitated the rapid discovery of mycoviruses. This has not only increased our understanding of mycoviral diversity within the fungal kingdom but has shed some light on the evolution and origins of these viruses. Two theories have been proposed. The ‘ancient coevolution theory’ posits that the relationship between mycoviruses and fungi is ancient and mirrors long-term co-evolution [3,5]. This is supported by the fact that mycoviruses lack an extracellular infection route, which complicates their transmission to other fungal species, and thus limits individual viruses to a single host [31]. Research by Neupane, et al. [32] also shows support for this theory, since the phylogenetic analysis of the RNA-dependent RNA polymerase (RdRp) sequences of mitoviruses from non-pathogenic arbuscular mycorrhizal fungi has revealed that they are highly conserved, and do not cluster with mitoviruses from other plant pathogenic fungi. The ‘plant virus theory’ on the other hand, suggests that mycoviruses originate from plant viruses, where they moved from the plant host to a fungus [3]. In support of this theory, studies have shown that mycoviruses often cluster with plant viruses on phylogenetic trees [33,34]. For example, Cryphonectria hypovirus 1–4 (CHV1-4), Fusarium graminearum virus 1 (FgV1), and Botrytis virus X are all related to plant potyviruses or potex-like viruses [5,35]. In addition, most mycoviruses lack a movement protein, and in some cases even a coat protein, which indicates that these viruses may have shed their non-essential genes to better adapt to their fungal hosts [31]. Natural cross-kingdom transfer of certain viruses between plants and fungi has also been demonstrated. Andika, et al. [36] provided evidence of this phenomenon by demonstrating the transmission of cucumber mosaic virus (CMV) from potato plants to Rhizoctonia solani through natural means. The transfer of plant viruses to fungi has also been shown by Cao et al. [37], where 11 different plant viruses were shown to transiently infect several plant pathogenic fungal genera, including Alternaria, Lecanicillium and Sarocladium. Cross-kingdom transmission has been proposed as a significant driver in the evolution of mycoviruses, given that the majority of mycovirus lineages have an ancestral link to plant viruses [38]. The opposite of this theory may also be true, where plant viruses may have originated from mycoviruses which moved from a fungus to a plant. For example, a recent study has demonstrated that mitoviral sequences can cross-transfer between Botrytis cinerea and cucumber plants [39]. Similarly, CMV has also been shown to move from R. solani to the plant host [36]. An alternative pathway for mycovirus transfer between fungi and plants, through the intermediary of a plant virus, has also been suggested. Bian, et al. [40] showed that CHV-1 was able to infect Nicotiana tabacum when co-inoculated with tobacco mosaic virus (TMV), a plant virus. Moreover, the findings suggest that CHV-1 may facilitate the accumulation of TMV in Fusarium graminearum, indicating a potential interaction between these viruses in a plant-fungal system. Due to a dearth of solid evidence and data, however, the exact origins of mycoviruses remains uncertain.

3. Diversity and Taxonomy

3.1. Fungal Host Range

Mycoviruses have been found across most of the major fungal taxonomic groups [1,5].
The majority of these mycoviruses associate with hosts from Ascomycota and Basidiomycota (collectively called Dikarya), as depicted in Table 1 and Table 2 [2]. However, this is most likely due to disproportionate sampling across the fungal kingdom. Recent studies have thus focused on the discovery of mycoviruses within the early diverging lineages of fungi and has led to the discovery of mycoviruses in Chytridiomycota, Zygomycota, Blastocladiomycota, and Neocallimastigomycota (Table 3) [2,41]. The ecological impacts of these viruses on their fungal counterparts, however, remains unknown and requires further investigation.

3.2. Mycoviral Taxa

Mycoviruses are currently classified into 23 families and one unclassified genus (Figure 1) by the ICTV according to their genome type and organization; https://talk.ictvonline.org/ (accessed on 27 March 2023). The majority of mycoviruses studied to date have dsRNA or ssRNA genomes, however, a few mycoviruses with ssDNA genomes have also been found [12,52,53,175,328].

3.2.1. Double-Stranded RNA (dsRNA) Mycoviruses

The genome characteristics of mycoviruses, including size and segmentation, exhibit variation across families and genera. Specifically, the genomes of dsRNA mycoviruses can range from non-segmented, as observed in Amalgaviridae and Totiviridae, to bisegmented in Megabirnaviridae, Partitiviridae, and Botybirnavirus, quadripartite in Quadriviridae, or multisegmented in Chrysoviridae, Polymycoviridae, and Spinareoviridae [5,337,338,339,340,341,342,343]. Furthermore, genome size among dsRNA mycoviruses is highly diverse, with some families containing mycoviruses with genomes as small as 3.0 kb and others having members with genomes as large as 29 kb [338,339]. Detailed information on genome length and segmentation of dsRNA mycoviruses is provided in Table 4.
In addition to differences in genome size and segmentation, the genome organization of mycoviruses also varies among families and genera. Although the presence of an RNA-dependent RNA polymerase (RdRp) domain is ubiquitous among all mycoviruses, its location within the genome may differ [2]. For instance, the RdRp domain in multisegmented viruses of the Chrysoviridae family is encoded on the first genome segment, while in non-segmented viruses of the Amalgaviridae family it is located on the 3′ proximal open reading frame (ORF) [337,342]. With the exception of members of the Botybirnavirus genus, all dsRNA mycoviruses encode a capsid protein (CP) [5,337,338,339,340,341,342]. Additionally, some mycoviruses contain domains encoding proteins with unknown functions, as well as specific domains specific to certain families, such as the phytoreo S7 domain in the Chrysoviridae, a methyltransferase (Mtf) and proline-alanine-serine-rich protein (PASrp) domain in Polymycoviridae, and the guanylyltransferase (Gtf) domain in Spinareoviridae [342,343,344]. Further details on the variation of genome organization among dsRNA mycovirus families can be found in Table 4.
Table 4. Genomic features and characteristics of dsRNA mycoviruses.
Table 4. Genomic features and characteristics of dsRNA mycoviruses.
Families and GeneraGenome Size and SegmentationGenome OrganizationExemplar Species
Family
Amalgaviridae
Mycoviral associated genera
Zybavirus
Non-segmented genomes around 3.5 kb in size.5′ proximal ORF encodes for a CP, followed by ribosomal frameshift signal and a 3′ proximal ORF encoding for an RdRp protein.Zygosaccharomyces bailii virus Z
Family
Chrysoviridae
Mycoviral associated genera
Alphachrysovirus
Betachrysovirus
Multisegmented genomes around 8.9 to 16.0 kb in size.
Alphachrysovirus: 3–4 separately encapsidated segments.
Betachrysovirus: 4–7 separately encapsidated genomes.
dsRNA 1: Encodes for an RdRp.
dsRNA 2: Encodes for a CP.
dsRNA 3: Encodes for phytoreo S7 domain in Alphachrysovirus, and an unknown protein in Betachrysovirus.
dsRNA 4–7: Encodes for a hypothetical protein with unknown function.
Alphachrysovirus:
Penicillium chrysogenum virus
Betachrysovirus:
Botryosphaeria dothidea chrysovirus 1
Family
Megabirnaviridae
Mycoviral associated genera
Megabirnavirus
Bisegmented genomes around 16 kb in size.dsRNA 1: 5′ proximal ORF encodes for a CP, followed by a ribosomal frameshift signal and a 3′ proximal ORF encoding for an RdRp.
dsRNA 2: Encodes for hypothetical protein with unknown function.
Rosellinia necatrix Megabirnavirus 1
Family
Partitiviridae
Mycoviral associated genera
Alphapartitivirus
Betapartitivirus
Gammapartivirus
Bisegmented genomes around 3 to 4.8 kb in size.dsRNA 1: Encodes for an RdRp.
dsRNA 2: Encodes for a CP.
Alphapartitivirus:
Rosellinia necatrix partitivirus 2
Betapartitivirus:
Ceratocystis resinifera virus 1
Gammapartivirus:
Aspergillus ochraceous virus
Family
Polymycoviridae
Mycoviral associated genera
Polymycovirus
Multisegmented genomes around 7.5 to 12.5 kb in size.dsRNA 1: Encodes for an RdRp.
dsRNA 2: Encodes for hypothetical protein with unknown function.
dsRNA 3: Encodes for a Mtf.
dsRNA 4: Encodes for a PASrp.
dsRNA 5–8: Encodes for hypothetical proteins with unknown functions.
Aspergillus fumigatus tetramycovirus 1
Family
Polymycoviridae
Mycoviral associated genera
Polymycovirus
Multisegmented genomes around 7.5 to 12.5 kb in size.dsRNA 1: Encodes for an RdRp.
dsRNA 2: Encodes for hypothetical protein with unknown function.
dsRNA 3: Encodes for a Mtf.
dsRNA 4: Encodes for a PASrp.
dsRNA 5–8: Encodes for hypothetical proteins with unknown functions.
Aspergillus fumigatus tetramycovirus 1
Family
Quadriviridae
Mycoviral associated genera
Quadrivirus
Quadripartite genomes around 3.5 to 5.0 kb in size.dsRNA 1: Encodes for a hypothetical protein with unknown function.
dsRNA 2 and 4: Encode for a CP.
dsRNA 3: Encodes for an RdRp.
Rosellinia necatrix quadrivirus 1
Family
Spinareoviridae
Mycoviral associated genera
Mycoreovirus
Multisegmented genomes (11–12 segments) around 23 to 29 kb in sizeSegments encodes for a single viral protein (VP1 to VP12)
VP 1: Encodes for an RdRp
VP 2: Encodes for CP
VP 3 or VP10: Encodes for a capping enzyme, (guanylyltransferase)
VP4 to VP12: Encodes for hypothetical proteins with unknown function
Mycoreovirus-1/Cp9B21
Family
Totiviridae
Mycoviral associated genera
Totivirus
Victorivirus
Non-segmented genomes around 4.6–7.0 kb in size5′ proximal ORF encodes for CP and the 3′ proximal ORF encodes for an RdRp
Totiviruses additionally encode for a ribosomal frameshift signal between the CP and RdRp
Some totiviruses may additionally produce satellite dsRNA which encode for killer proteins
Totivirus:
Saccharomyces cerevisiae virus L-A
Victorivirus:
Helminthosporium victoriae virus 190S
Family
Unclassified
Mycoviral associated genera
Botybirnavirus
Bisegmented genomes around 5.7 to 6.3 kb in sizedsRNA 1: Encodes for an RdRp
dsRNA 2: Encodes for a hypothetical protein with unknown function
Bipolaris maydis botybirnavirus 1
Compiled using data from the International Committee on Taxonomy of Viruses (ICTV): https://ictv.global/taxonomy/ (accessed on 20 March 2023), Nibert, Ghabrial, Maiss, Lesker, Vainio, Jiang and Suzuki [33], Krupovic, et al. [337], Ghabrial, Castón, Jiang, Nibert and Suzuki [5], Lin, et al. [345], and Li, et al. [346].

3.2.2. Single Stranded RNA (ssRNA) Mycoviruses

Positive (+) Sense

Mycoviruses belonging to families with positive-sense single-stranded RNA (+ssRNA) genomes display considerable diversity in terms of genome size, with genomes ranging from approximately 2.0 kb to 17.6 kb [172,347]. The majority of +ssRNA mycoviruses have non-segmented genomes, however, the recently classified family Hadakaviridae consists of members with 10 to 11 genome segments [7]. While most mycoviruses replicate within the cytoplasm of their host, members of the Mitoviridae family demonstrate a unique replication pattern, taking place in the mitochondria of its fungal host [172]. More detail on the diversity of genome length and segmentation among +ssRNA mycovirus families and genera is presented in Table 5.
The presence of an RdRp domain is a characteristic feature of all +ssRNA mycoviruses. The genome of these viruses can include different protein domains, such as viral helicases (Hel), methyl transferases (Mtf), glycosyl transferases (Gtf), capsular polysaccharide synthases (CPS), phytoreo S7 domains, capsid proteins (CP), and proteases [347,348,349,350,351]. The location of the RdRp domain within the genome can vary among different mycovirus families, as can the specific protein domains present in their genomes. Further information on the variation of genome organization among +ssRNA mycovirus families can be obtained from Table 5.

Negative (−) Sense

Negative sense ssRNA mycoviruses belong to a single family, namely Mymonaviridae, which is characterized by filamentous, enveloped viruses with linear genomes that are around 6 to 10 kbp [352]. The genomes of −ssRNA viruses encode for one or more proteins, two of which are known, namely an RdRp protein and a nucleoprotein (NP) which encloses the viral genome [230,279,352,353]. The functions of all other ORF encoded proteins are still uncertain [352]. Further details on genome organization and genera within Mymonaviridae is given in Table 6.

Reverse Transcribing (RT)

Concerning reverse transcribing ssRNA mycoviruses, two families are currently recognized: Metaviridae and Pseudoviridae. Mycoviruses from the Metaviridae family are characterized by genomes ranging in length from 3 to 15 kb, while members of the Pseudoviridae family are generally shorter, ranging from 4 to 9 kb in length [354,355]. The replication process of these viruses involves reverse transcription within intracellular virus-like particles (VLP) to generate complementary DNA (cDNA), which is then integrated into the host chromosome through the action of an integrase protein [354,355].
The genome of RT ssRNA mycoviruses typically encodes for capsid (CP) or nucleocapsid proteins (NC), which are located on the gag gene, as well as protease (PR), reverse transcriptase (RT), integrase (INT), and ribonuclease H domains (RH), which are located on the pol gene [354,355,356]. The key difference between the two families lies in the fact that members of the Pseudoviridae family may encode the gag and pol proteins on separate open reading frames (ORFs), while members of the Metaviridae family encode these genes on a single ORF [354,355]. Further information on the different genera and characteristics of RT ssRNA viral families can be found in Table 7.

3.2.3. Single-Stranded DNA (ssDNA) Mycoviruses

Mycoviruses with DNA genomes are rare compared to RNA viruses. There is currently only one recognized family containing ssDNA mycoviruses, namely Genomoviridae. This family encompasses two genera: Gemycircularvirus (represented by Sclerotinia sclerotiorum hypovirulence-associated DNA virus 1, SsHADV-1) and Gemytripvirus (represented by Fusarium graminearum gemytripvirus 1, FgGMTV1) [12,175]. These mycoviruses exhibit small genome sizes, which range from 1.3 kb to 2.4 kb [14,175]. The genomes of SsHADV-1 and FgGMTV1 encode for a replicase protein (REP) and a capsid protein (CP), with the former encoded on a single circular ssDNA genome, and the latter encoded on three circular ssDNA components (DNA-A, DNA-B, and DNA-C). ssDNA mycoviruses are not exclusively found in Genomoviridae, however. For example, geminivirus-like ssDNA has also been discovered in Macrophomina phaseolina (M. phaseolina DNA virus; MpDV) and Mucor racemosus (M. racemosus DNA virus; MrDV) [53]. Additionally, a novel ssDNA mycovirus, Gigaspora circovirus A has also been discovered and belongs to the Circoviridae family but has not yet been assigned to the genus [52]. Unlike other circoviruses which encode for both a REP and CP, Gigaspora circovirus A encodes for a REP only [52]. Although not officially recognized by the ICTV, Gigaspora circovirus A might represent the first member of Circoviridae which infects a fungal host [14,175]. Table 8 summarizes the properties of ssDNA mycoviruses which are currently recognized by ICTV.

4. Effect of Mycoviruses on Fungi

By being obligate intracellular parasites mycoviruses depend on host genes and gene products for their replication, which will in turn affect how the host phenotype manifests [6]. The majority of mycoviruses described so far have had no observed effect on their fungal hosts, although some can alter their host’s phenotypic traits [5]. These mycoviruses may change the virulence of their fungal hosts, often by altering the transcriptome profiles of their fungal hosts by interfering with protein–protein interactions and silencing antiviral proteins [1,6]. It should be noted, however, that these associations are sometimes complex. Several environmental factors, including the growth media [158], temperature [357], other mycoviruses [358], and the fungal plant or animal host [186] may also play a role in the outcome of the fungal host phenotype.

4.1. Host–Virus Arms Race

Mycoviral evasion of host defenses and underlying mechanisms of antiviral mechanisms are commonly studied using Cryphonectria parasitica, the model organism for mycoviral research [21,359,360], and more recently Neurospora crassa [198].
C. parasitica can defend against viral infection by use of an RNA-mediated gene regulation mechanism known as RNA silencing or RNA interference (RNAi) which commonly involves the use of endoribonucleases known as Dicer-like proteins (DCLs), and also Argonaute-like proteins (AGLs) [361,362,363]. DCL and AGL proteins form part of the RNA-induced silencing complex (RISC), which recognizes and cleaves the dsRNA genomes, or replicative intermediates, of infecting mycoviruses [6]. An in-depth review of RNA silencing mechanisms is covered by Zhao and Guo [364].
Mycoviruses may also suppress RNA silencing. Through proteins such as p29 and p27, Cryphonectria hypovirus 1 (CHV-1) and Cryphonectria hypovirus 2 (CHV-2) suppress this pathway in C. parasitica, allowing enhanced replication and altered symptoms in the host [365,366]. Similarly, Rosellinia necatrix mycoreovirus 3 (RnMYRV-3) produces S10, which also helps suppress silencing in Rosellinia necatrix [367]. In N. crassa, RISC-associated proteins are upregulated after infection with Neurospora crassa fusarivurs 1 (NcFV1), but AGL proteins are downregulated post-transcriptionally [198]. On the other hand, some mycoviruses, such as Talaromyces marneffei partitivirus-1 (TmPV- 1), downregulate the mRNA levels of DCL and AGL proteins in the fungal host [110].
Interestingly, some types of fungi do not have an RNAi pathway, rendering them highly susceptible to mycovirus infection. For example, the persistence and replication of MsMV-1, a mycovirus of Malassezia sympodialis (a yeast linked to common skin disorders, pancreatic cancer, and Chron’s disease), is thought to be due to the absence of an RNAi pathway within this organism [311,368,369,370].

4.2. Hypervirulence, Hypovirulence and the Effect of Mycoviruses on the Pathogenesis of Fungi

Mycoviruses that induce hypervirulence increase the virulence of the fungal host [6]. Despite its perceived negative effects, hypervirulence may also have desirable traits in plant pathogenic fungi. For example, amycovirus known as Leptosphaeria biglobosa quadrivirus 1 (LbQV-1) induces hypervirulence in Leptosphaeria biglobosa [77]. The infected strains of L. biglobosa confer systemic resistance and protect the host plant, Brassica napus (oilseed rape or canola), from a closely related and more aggressive fungus called Leptosphaeria maculans [77]. Some mycoviruses also confer hypervirulence to entomopathogenic fungi from the genera Metarhizium and Beauveria, which are routinely used as biocontrol agents against a few arthropod pests [371]. In these circumstances, hypervirulence is a more desirable trait as it may increase the effectiveness of fungicides that are currently on the market.
Mycoviruses that induce hypovirulence, adversely affect the virulence, morphology, sporulation, growth rate, and pigmentation of their fungal hosts [6,15,16]. Although the precise nature and molecular pathways by which these viruses cause hypovirulence are still largely unclear, they are known to exert some of their effects through virus–host protein–protein interactions [6]. Mycovirus-induced hypovirulence of C. parasitica can be attributed to the alteration of numerous signal transduction pathways, including those important for virulence [372,373,374]. For example, the reduction of pigmentation, sporulation, and laccase accumulation in C. parasitica is linked to the CHV-1 papain-like protease p27 [375]. CHV-1 infected C. parasitica additionally upregulates ATG8, a homolog of ubiquitin-like yeast autophagy protein, since it is essential for replicating this virus [376]. In contrast, CHV-1 also differentially regulates proteins that prevent the expression of viral RNA, such as DNA methyltransferases, which in turn results in retarded growth and aberrant colony morphology of the host [377].
Mycoviruses that induce hypovirulence have been discovered in various plant pathogenic fungi, including several Fusarium spp. [349,378,379], the white root rot fungus R. necatrix [213,215,380], white mold fungus S. sclerotiorum [12,16,381], rice blast fungus Magnaporthe oryzae [382], and the grey mold rot fungi Botrytis cinerea [39,383,384] and Alternaria spp. [82,385]. Hypovirulence-inducing mycoviruses are of immense interest to plant pathologists, due to their potential use as biocontrol agents of plant pathogenic fungi.
The precise impact of mycoviruses on pathogenic fungi affecting human health remains largely unknown, yet a handful of studies have shed some light on these relationships. For example, infection with MsMV-1 appears to result in significant transcriptional rewiring in M. sympodiali, causing upregulation of transcriptional factors and ribosomal genes, while simultaneously repressing genes responsible for cellular metabolism [311]. Interestingly, the MsMV-1 putative mycoviral toxin also elicits an immune response in macrophages and augments the ability of infected isolates to colonize murine skin [311]. This suggests that the mycovirus may play a role in the pathogenicity of the host fungus. Similarly, recent investigations have discovered the presence of antibodies against a certain strain of mycovirus-infected Aspergillus flavus (MCAF) in the plasma of patients that were in remission from acute lymphoblastic leukemia (ALL) [386]. Within this study, the exposure of mononuclear blood cells from ALL patients in remission to the supernatant of a mycovirus-containing Aspergillus flavus resulted in a significant reappearance of cell surface and genetic markers consistent with this disease [386]. The authors, therefore, hypothesize that exposure to MCAF may contribute to the development of ALL [386]. These studies show that the role of mycoviruses in diseases caused by human pathogenic fungi, with and without their hosts, needs to be further explored.

4.3. The ‘Killer Phenotype’ in Yeasts

Mycoviruses do not exclusively infect filamentous fungi but have also been associated with various species of yeast [264,268]. Some induce the ‘yeast killer phenotype’, eliminating competing yeasts and providing a competitive advantage to the host [387]. The killer phenotype in the model organism S. saccharomyces is normally determined by two co-infecting totiviruses with a mutualistic relationship, namely the helper LA and satellite M virus [388]. ScV (Saccharomyces Cerevisiae virus) -L-A encodes essential proteins for replication and transcription, while satellite M encodes for a preprotoxin which provides immunity to the host and kills off non-infected cells when processed into toxin [387]. The toxins generated by this satellite virus vary among strains and can induce cell death through different mechanisms. These include the disruption of cytoplasmic membrane function and the formation of lethal ion channels (K1 and K2), as well as cell cycle arrest in the G1 or early S phase (K28) [389,390,391].
Industrial yeast strains with virally encoded killer systems are highly sought after for their ability to restrain spoilage microorganisms and preserve the quality of food products and beverages [392,393,394]. For example, Ustilago maydis, Kluyveromyces wickerhamii, Pichia anomala, and Pichia membranifaciens produce virally encoded toxins that have antifungal activity against the wine spoilage yeast, Brettanomyces bruxellensis [395,396]. They have also been discovered in Zygosaccharomyces bailii, where they encode for the toxin zygocin, which has broad antifungal activity and has the potential as an antimycotic drug [397].

5. Mycoviruses as Biocontrol Agents

Among the best-known applications of a mycovirus in the field is CHV-1 against the chestnut blight pathogen, C. parasitica [398,399]. Under typical conditions, C. parasitica infections result in the appearance of cankers on the stems and branches of susceptible trees, which destroy the cambium tissue and ultimately result in tree death [400]. CHV-1 infected strains, on the other hand, have reduced virulence, resulting in superficial cankers which eventually stop growing and become passive [400]. Mycovirus-based biocontrol of chestnut blight has proven to be a great success in Europe, largely due to the low genetic diversity among fungal strains [400]. In regions where naturally hypovirulent strains of C. parasitica were present, CHV-1 effectively spread without intervention [401]. However, in areas with little to no natural hypovirulence, CHV-1 could also be artificially introduced by treating bark cankers with hypovirulent C. parasitica [402,403]. Research into hypovirulence-associated mycoviruses in other plant pathogenic fungi has since been prompted by the success of hypovirus-mediated hypovirulence in C. parasitica.
The Sclerotinia sclerotiorum hypovirulence-associated DNA virus 1 (SsHADV1) is infectious as purified particles, and can directly infect the hyphae of S. sclerotiorum [404]. Researchers have thus developed an aerial spray capable of killing S. sclerotiorum on infected rapeseed plants in the field, by using hyphal fragments from an infected strain of the fungus [404]. Sclerotinia sclerotiorum partitivirus 1 (SsPV1), another mycovirus of S. sclerotiorum, appears to spread through hyphal contact to different strains regardless of vegetative incompatibility [20]. This suggests that SsPV1 also has potential as a biocontrol agent against S. sclerotiorum in the field [20].
While the focus of mycoviral research has predominantly centered around edible mushrooms and plant pathogenic fungi, an emerging area of investigation pertains to the search for mycoviruses with potential therapeutic benefits for human health [107,357,405].

5.1. Limitations of Biocontrol: The Role of Mycovirus Transmission

Despite their immense potential, the use of mycoviruses as biocontrol agents of plant pathogenic fungi is complicated by their mode of transmission. Mycoviruses lack an extracellular route of transmission [3]. Instead, they can only spread intracellularly, through hyphal anastomosis (horizontal transmission) or sporulation (vertical transmission) [5].

5.1.1. Horizontal Transmission

One of the greatest barriers to the successful spread of mycoviruses pertains to hyphal anastomosis. Hyphal anastomosis occurs when specialized hyphae from the same fungus, or hyphae from different fungi fuse and exchange cytoplasmic content, which includes any associated mycoviruses [406]. However, in order for the hyphae from different fungal strains or species to fuse, they need to be vegetatively compatible [3]. Vegetative compatibility is determined by the fungal vegetative incompatibility genes (vic genes), which will trigger programmed cell death (PCD) when contact between incompatible fungi occurs [407]. Thus, mycoviruses cannot be transmitted from a hypovirulent fungal strain to a target fungal strain if they are vegetatively incompatible [21]. For example, CHV1 has been effective against C. parasitica in Europe, but not in America, where there is greater diversity of vic groups between fungal strains [401,408]. Similarly, although S. sclerotiorum harbours a diverse range of mycoviruses, their use in the field is restricted due to the high diversity of vic loci among different strains which may also be very complicated under field conditions [409,410]. Research is underway to resolve the issue of vegetative incompatibility as it relates to mycovirus transmission.
Many methods and techniques have been developed in order to study vegetatively compatibility systems in fungi and to find ways to overcome the vic system. For example, vic genes related to five to six loci in C. parasitica have been linked to vegetative incompatibility and virus transmission in one study [21]. The disruption of these genes then allowed for the development of a super donor strain, which allowed the spread of mycoviruses between incompatible strains [411]. This approach will not be feasible for all fungi, however, as some plant pathogenic fungi may have more complicated vic systems that result in a high vic diversity [16,228]. Chemical compounds have been used to enhance viral transmission and to prevent programmed cell death (PCD), which is caused by vegetative incompatibility between fungi [412]. When vegetatively incompatible strains of R. necatrix were cultured together on a medium supplemented with zinc, hyphal anastomosis improved and mycoviruses could be transmitted to isolates of different VCGs [413].
While vegetative incompatibility prevents transmission of mycoviruses in most cases, other factors also often play a role. In situ inoculation on chestnut wood increased transmission efficiency of CHV1 between vegetatively incompatible strains [414]. In another study, two vegetatively incompatible strains of R. necatrix, one of which contained a mycovirus, were inoculated on apple trees and later found to harbor the same mycovirus [415]. This may be due to several factors, including different environmental conditions, and a weakened vegetative incompatibility response due to environmental microorganisms or the host plant itself [407]. Research has also indicated that horizontal transmission of mycoviruses between fungal and plant hosts may occur in cases of co-infection with plant viruses [40]. In such instances, the replication of these viruses in both hosts is facilitated. This phenomenon has been linked to suppressing antiviral mechanisms in both the plant and fungus [40]. For example, the mycovirus CHV-1 produces a protein called p29 that downregulates components of the antiviral RNAi system in fungal hosts, thereby promoting virus accumulation [365]. However, research by Bian et al. [39] has shown that this protein has limited functionality in certain plant hosts, such as N. tabacum. TMV on the other hand, encodes for a replicase that interferes with the antiviral response in the plant host, thereby enhancing CHV-1 accumulation in the plant [40]. Additionally, TMV produces a cell-to-cell movement protein (MP) that is typically absent in mycoviruses, thus enabling the dissemination of CHV-1 throughout the N. tabacum [365]. This may then enhance the ability of the mycovirus to access other fungi that may have established themselves in the same plant [40]. Conversely, CHV-1 inhibits the fungal antiviral defense mechanism, which typically acts to eliminate the plant virus [40]. This, in turn, permits the accumulation of TMV within the fungal host, specifically F. graminearum, as demonstrated in this study [40].
Different viral strains also play a role in transmission efficiency, where researchers have found that strains with higher virulence have higher transmissibility [416].
Interestingly, several mycoviruses are known to either influence the host vic system to transmit between vegetatively incompatible strains or infect the host directly as infectious particles [188,234,404,417]. CHV-1 for example, has been found to downregulate genes that are involved in programmed cell death (PCD), which occurs after vegetatively incompatible strains interact, thus allowing for transmission between incompatible strains [417,418]. Co-infection of mycoviruses can also result in the transmission of viruses between vegetatively incompatible fungi. One study has demonstrated that Sclerotinia sclerotiorum mycoreovirus 4 (SsMYR4) downregulates cellular activities and pathways associated with vegetative incompatibility mediated PCD [419]. This in turn facilitated the horizontal transmission of other hypovirulent co-infecting viruses [419]. Mycoviruses with these traits hold immense potential as biocontrol agents; however, more research is required to fully understand the mechanisms behind these phenomena.
Regarding their clinical application, mycoviruses that induce hypovirulence would require administration through delivery methods such as injection or topical application to the target fungus of an infected patient [107]. The lack of extracellular replication in mycoviruses is caused by the impenetrable fungal cell wall acting as a barrier against mycovirus uptake, thus limiting their potential as therapeutic agents against human pathogenic fungi [420]. Hyphal anastomosis is a possible alternative, as noted previously, various challenges preclude its practical use in human patients [107]. A promising avenue for the development of therapeutic interventions for humans has emerged with the discovery of ssDNA mycoviruses that are capable of extracellular transmission in S. sclerotiorum [107,404]. It is thus plausible that ssDNA viruses may represent the most viable candidates for therapeutic applications in humans. Because mycoviruses lack an extracellular route of infection, transfection, or transformation using full-length viral cDNA clones, purified virus particles, and in vitro RNA transcripts are usually used to transmit mycoviruses between incompatible fungi in the laboratory [421,422,423]. The development of such clones is complex, however, especially in the case of multisegmented mycoviruses [407]. Nevertheless, studies have demonstrated that encapsulated mycoviruses can be transformed into fungal protoplasts using polyethylene glycol-mediated protocols, which is now the standard approach for transmission in a laboratory setting [12,424,425]. Mycoviruses can also be transmitted between incompatible fungal strains through protoplast fusion, which has the advantage of allowing transmission of both encapsulated and unencapsulated viruses [22]. Although these methods normally study host and viral factors involved in viral replication or symptom induction, they have also been used to expand the host range for some mycoviruses [22].
Vectors, such as insects or parasites that transmit mycoviruses between fungi, may be used to overcome barriers of vegetative compatibility and rapidly help mycoviruses establish populations in the field [20]. For example, the transmission of SsHADV-1 by frugivorous insects to other vegetatively incompatible strains has been observed under laboratory conditions, for example [20]. However, producing and dispersing such vectors in the field is impractical [20]. Hence, it is more feasible to exploit a naturally occurring vector, such as mycoparasites, which can transmit the mycovirus via hyphal parasitization [20]. A hypovirulence-associated mycovirus can thus be introduced into the mycoparasite using transfection techniques or dual culturing and then used to infect a host fungus [20].
Most mycoviral research regarding biocontrol focuses on hypovirulence-inducing viruses, but some researchers have also exploited the “killer phenotypes” of dimorphic fungi to confer resistance to plant hosts. For example, one study has shown that transgenic expression of the viral KP4 killer toxin from U. maydis into Swiss wheat confers resistance to these crops against U. maydis and related hosts [426]. These killer systems are not generally present in filamentous fungi, but similar strategies could also be explored for plant pathogenic yeasts or dimorphic fungi.

5.1.2. Vertical Transmission

For mycoviruses to be considered effective biocontrol agents, they must not only exhibit hypovirulence and have the ability to transmit to uninfected fungi through hyphal anastomosis, but they must also demonstrate efficient transmission to the fungal progeny. This requirement is crucial for the long-term efficacy of mycovirus-based biocontrol strategies. The transmission of some mycoviruses to the fungal progeny occurs primarily through sporulation, which can be sexual or asexual [1,3,15]. Transmission rates, however, vary greatly between fungus-virus combinations and between different spore types (asexual vs. sexual) [15]. For example, one study has demonstrated that the transmission of CHV1, 2, 3, and 4 to the ascospore progeny of C. parasitica is ineffective as infection results in a loss of female fertility [118]. In contrast to this, other studies have shown that mycoviruses can be transmitted to the ascospore progeny of C. parasitica strains which are infected by reoviruses (Mycoreovirus 1 and 2) or a mitovirus (Cryphonectria mitovirus-1) [118,427]. Vertical transmission through asexual spores or other asexual structures such as sclerotia are commonly observed in mycoviruses [228,379,428,429]. Disseminating mycoviruses into asexual spores allows the spread of these viruses within their host to longer distances than permitted by horizontal transmission alone and allows mycoviruses to persist within the fungal population [414]. Mycovirus-based biocontrol strategies may face challenges with vertical transmission through asexual fragments. In some cases, hypovirulence has been observed to be linked with the disruption of conidia, resulting in lower transmission rates. For instance, CHV-2 induces substantial hypovirulence in C. parasitica, but has limited transmission to conidia with only a 2–5% transmission rate, leading to its restricted geographical distribution [430]. It is important to note that this scenario may not be universal. According to a study by Lee, et al. [431], four mycoviruses, which reduce the virulence of Fusarium graminearum, exhibit more efficient transmission to conidia than those that cause symptomless infections. Another issue related to the transmission of fungal pathogens through asexual spores is that some fungal species either do not produce conidia or the conidia play a minimal role in their life cycle [407].
In general, the transmission of mycoviruses to sexual spores (ascospores, basidiospores, etc.) is believed to be less prevalent than the transmission of mycoviruses to asexual spores (conidia) [20]. However, studies now indicate that this transmission mode may be more common than previously thought [118,232,432,433]. The mechanisms behind mycovirus transmission via sexual spores remain unclear, but it is believed that this could represent a potential pathway for exchanging mycoviruses among different vic groups [407]. Hence, an increased focus on investigating these mechanisms could provide new avenues for controlling mycovirus transmission and boosting the potential of biological control efforts.

6. The Detection of Mycoviruses

6.1. In Vitro Based Detection

In the past, mycoviruses were primarily detected by the use of a culture-based approach. The basis for culture-based detection is rooted in the observation that most mycoviruses have dsRNA genomes, or a dsRNA replicative intermediate, which are not generated by the host [5,15]. The standard method used to detect mycoviruses with RNA genomes involves purifying dsRNA from total fungal RNA extracts, by using cellulose chromatography or other column-based approaches [434,435]. While these approaches are usually rapid and inexpensive, they suffer from several disadvantages [2]. Using an in vitro-based approach may lead to overestimation of dsRNA or ssRNA levels since dsRNA enrichment protocols are commonly used [3]. Moreover, these techniques strengthen the idea that mycovirus genomes are primarily composed of RNA and completely disregard viruses that may be composed of DNA [2]. Purified dsRNA is also normally visualized using agarose gel electrophoresis, which may result in false negatives in cases where there are only low-titer infections.
Culture-based methods are also used to evaluate the effect of mycoviral infection on the fungal host. Infected cultures may show decreased growth and sporulation, as well as alteration in pigmentation and morphology [6]. For example, CHV-1-infected C. parasitica are deficient in the characteristic orange pigmentation of non-infected strains. In contrast, infected S. sclerotiorum strains show abnormal colony morphology and show smaller and fewer sclerotia [6]. However, this is not always the case. Some mycoviruses that induce hypovirulence in vitro, may induce hypervirulence in planta [84]. Consequently, dsRNA profiling is usually used alongside in silico methods to identify and characterize mycoviruses [2,435].

6.2. In Silico Based Detection

Thanks to the development of new technologies, it has become easier to detect mycoviruses and gain insight into their structure and biology [2,20,436]. By interrogating the metatranscriptomic datasets of fungal hosts for sequences showing homology to mycoviral proteins, the in silico approach can identify any associated putative mycoviruses. One such protein is the RdRp, which is essential for viral transcription and replication, and is thus ubiquitous in RNA viruses [2]. Other proteins which are also found in the genomes of some, but not all mycoviruses, are viral helicases, glycosyl transferases, methyl transferases, and capsid proteins [211,437,438]. Because the in silico approach to detection does not make use of agarose gel electrophoresis, it is more sensitive to low titer mycoviral infections than in vitro-based approaches [2]. The greater accessibility of fungal transcriptomic datasets on open-access platforms like the NCBI sequencing reads archive (SRA), as a result of decreased sequencing costs, is another benefit of in-silico-based approaches [61]. Researchers can now examine publicly available fungal datasets for the presence of these viruses, further elucidating the diversity and prevalence of mycoviruses in under-researched fungal groups.
Some studies now use approaches such as genome-wide linkage analysis to understand the underlying mechanisms behind the effects that mycoviruses have on their fungal hosts [431]. For instance, RNA-seq-based genome-wide expression analyses showed distinct expression patterns in response to infection by four phylogenetically different mycoviruses in F. graminearum (FgV1-4) [431]. Even though these mycoviruses all showed changes in transcriptome expression, only FgV1 and FgV2 caused observable changes in the host phenotype [431]. Mycoviruses are dependent on many host factors, as well as pathways and processes related to metabolism, transport, RNA processing, and signaling, and not all of these will result in a phenotypic change in the host [3]. Mycoviruses and their hosts often interact in complex ways, so more detailed research is needed to better understand these interactions.

7. Conclusions

Mycoviruses are ubiquitous within the fungal kingdom. Recent studies have shown that they are associated with most of the major fungal taxa. Yet, the number of mycoviruses that have been fully characterized is low in comparison to economically important plant and animal viruses. This is most likely due to their cryptic nature, which means that it is often difficult to distinguish whether a fungal host is infected by empirical observation. Mycoviruses have been commonly detected with in vitro-based approaches, however, these are known to suffer from several disadvantages. The rapid advances in high throughput sequencing however, particularly RNA sequencing, which has become mainstream, has led to an exponential increase in the number of mycoviruses that have been discovered. Fungal transcriptomes can now be mined for mycoviral sequences, and this allows for the study of the complex interactions between fungal hosts and mycoviruses. Unlike most other viruses, mycoviruses do not always elicit measurable changes to host phenotypes. It is now clear that some mycoviruses cause hypovirulence and may be potential biocontrol agents of plant pathogenic fungi, which makes their discovery and characterization in such hosts even more important. There are, however, still numerous challenges that need to be addressed before their widespread use. Research is ongoing in pursuing the use of mycoviruses as biocontrol agents, and it is conceivable that these challenges will be overcome in the future.

Author Contributions

Writing—original draft preparation, B.H.; writing—review and editing, E.S., B.W. and D.R.; supervision, B.W. and D.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of South Africa, grant number 98353.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data was created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghabrial, S.A.; Suzuki, N. Viruses of Plant Pathogenic Fungi. Annu. Rev. Phytopathol. 2009, 47, 353–384. [Google Scholar] [CrossRef] [PubMed]
  2. Myers, J.; Bonds, A.; Clemons, R.; Thapa, N.; Simmons, D.; Carter-House, D.; Ortanez, J.; Liu, P.; Miralles-Durán, A.; Desirò, A. Survey of early-diverging lineages of fungi reveals abundant and diverse mycoviruses. Mbio 2020, 11, e02027-20. [Google Scholar] [CrossRef] [PubMed]
  3. Son, M.; Yu, J.; Kim, K.-H. Five Questions about Mycoviruses. PLoS Pathog. 2015, 11, e1005172. [Google Scholar] [CrossRef] [PubMed][Green Version]
  4. Walker, P.J.; Siddell, S.G.; Lefkowitz, E.J.; Mushegian, A.R.; Adriaenssens, E.M.; Dempsey, D.M.; Dutilh, B.E.; Harrach, B.; Harrison, R.L.; Hendrickson, R.C. Changes to virus taxonomy and the Statutes ratified by the International Committee on Taxonomy of Viruses (2020). Arch. Virol. 2020, 165, 2737–2748. [Google Scholar] [CrossRef]
  5. Ghabrial, S.A.; Castón, J.R.; Jiang, D.; Nibert, M.L.; Suzuki, N. 50-plus years of fungal viruses. Virology 2015, 479, 356–368. [Google Scholar] [CrossRef][Green Version]
  6. Kotta-Loizou, I. Mycoviruses and their role in fungal pathogenesis. Curr. Opin. Microbiol. 2021, 63, 10–18. [Google Scholar] [CrossRef]
  7. Sato, Y.; Turina, M.; Chiba, S.; Okada, R.; Bhatti, M.F.; Kotta-Loizou, I.; Coutts, R.H.; Kondo, H.; Sabanadzovic, S.; Suzuki, N. ICTV Virus Taxonomy Profile: Hadakaviridae 2023. J. Gen. Virol. 2023, 104, 001820. [Google Scholar] [CrossRef]
  8. Sato, Y.; Das, S.; Velasco, L.; Turina, M.; Osaki, H.; Kotta-Loizou, I.; Coutts, R.H.; Kondo, H.; Sabanadzovic, S.; Suzuki, N. ICTV Virus Taxonomy Profile: Yadokariviridae 2023. J. Gen. Virol. 2023, 104, 001826. [Google Scholar] [CrossRef]
  9. Lin, Y.-H.; Fujita, M.; Chiba, S.; Hyodo, K.; Andika, I.B.; Suzuki, N.; Kondo, H. Two novel fungal negative-strand RNA viruses related to mymonaviruses and phenuiviruses in the shiitake mushroom (Lentinula edodes). Virology 2019, 533, 125–136. [Google Scholar] [CrossRef]
  10. Guo, M.; Shen, G.; Wang, J.; Liu, M.; Bian, Y.; Xu, Z. Mycoviral diversity and characteristics of a negative-stranded RNA virus LeNSRV1 in the edible mushroom Lentinula edodes. Virology 2021, 555, 89–101. [Google Scholar] [CrossRef]
  11. Wang, L.; He, H.; Wang, S.; Chen, X.; Qiu, D.; Kondo, H.; Guo, L. Evidence for a novel negative-stranded RNA mycovirus isolated from the plant pathogenic fungus Fusarium graminearum. Virology 2018, 518, 232–240. [Google Scholar] [CrossRef]
  12. Yu, X.; Li, B.; Fu, Y.; Jiang, D.; Ghabrial, S.A.; Li, G.; Peng, Y.; Xie, J.; Cheng, J.; Huang, J.; et al. A geminivirus-related DNA mycovirus that confers hypovirulence to a plant pathogenic fungus. Proc. Natl. Acad. Sci. USA 2010, 107, 8387–8392. [Google Scholar] [CrossRef][Green Version]
  13. Krupovic, M.; Ghabrial, S.A.; Jiang, D.; Varsani, A. Genomoviridae: A new family of widespread single-stranded DNA viruses. Arch. Virol. 2016, 161, 2633–2643. [Google Scholar] [CrossRef][Green Version]
  14. Varsani, A.; Krupovic, M. Family Genomoviridae: 2021 taxonomy update. Arch. Virol. 2021, 166, 2911–2926. [Google Scholar] [CrossRef]
  15. Pearson, M.N.; Beever, R.E.; Boine, B.; Arthur, K. Mycoviruses of filamentous fungi and their relevance to plant pathology. Mol. Plant Pathol. 2009, 10, 115–128. [Google Scholar] [CrossRef]
  16. Jiang, D.; Fu, Y.; Guoqing, L.; Ghabrial, S.A. Viruses of the plant pathogenic fungus Sclerotinia sclerotiorum. In Advances in Virus Research; Academic Press: New York, NY, USA, 2013; Volume 86, pp. 215–248. [Google Scholar]
  17. Hao, F.; Ding, T.; Wu, M.; Zhang, J.; Yang, L.; Chen, W.; Li, G. Two Novel Hypovirulence-Associated Mycoviruses in the Phytopathogenic Fungus Botrytis cinerea: Molecular Characterization and Suppression of Infection Cushion Formation. Viruses 2018, 10, 254. [Google Scholar] [CrossRef][Green Version]
  18. Wu, M.; Jin, F.; Zhang, J.; Yang, L.; Jiang, D.; Li, G. Characterization of a Novel Bipartite Double-Stranded RNA Mycovirus Conferring Hypovirulence in the Phytopathogenic Fungus Botrytis porri. Virol. J 2012, 86, 6605–6619. [Google Scholar] [CrossRef][Green Version]
  19. Nuss, D.L. Hypovirulence: Mycoviruses at the fungal–plant interface. Nat. Rev. Microbiol. 2005, 3, 632–642. [Google Scholar] [CrossRef]
  20. Xie, J.; Jiang, D. New Insights into Mycoviruses and Exploration for the Biological Control of Crop Fungal Diseases. Annu. Rev. Phytopathol. 2014, 52, 45–68. [Google Scholar] [CrossRef][Green Version]
  21. Choi, G.H.; Dawe, A.L.; Churbanov, A.; Smith, M.L.; Milgroom, M.G.; Nuss, D.L. Molecular Characterization of Vegetative Incompatibility Genes That Restrict Hypovirus Transmission in the Chestnut Blight Fungus Cryphonectria parasitica. Genetics 2012, 190, 113–127. [Google Scholar] [CrossRef][Green Version]
  22. Lee, K.-M.; Yu, J.; Son, M.; Lee, Y.-W.; Kim, K.-H. Transmission of Fusarium boothii Mycovirus via Protoplast Fusion Causes Hypovirulence in Other Phytopathogenic Fungi. PLoS ONE 2011, 6, e21629. [Google Scholar] [CrossRef] [PubMed]
  23. Wagemans, J.; Holtappels, D.; Vainio, E.; Rabiey, M.; Marzachì, C.; Herrero, S.; Ravanbakhsh, M.; Tebbe, C.C.; Ogliastro, M.; Ayllón, M.A.; et al. Going Viral: Virus-Based Biological Control Agents for Plant Protection. Annu. Rev. Phytopathol. 2022, 60, 21–42. [Google Scholar] [CrossRef] [PubMed]
  24. Sinden, J.; Hauser, E. Report on two new mushroom diseases. Mushroom Sci. 1950, 1, 96–100. [Google Scholar]
  25. Hollings, M. Viruses Associated with A Die-Back Disease of Cultivated Mushroom. Nature 1962, 196, 962–965. [Google Scholar] [CrossRef]
  26. Ellis, L.; Kleinschmidt, W. Virus-like particles of a fraction of statolon, a mould product. Nature 1967, 215, 649–650. [Google Scholar] [CrossRef]
  27. Banks, G.; Buck, K.; Chain, E.; Himmelweit, F.; Marks, J.; Tyler, J.; Hollings, M.; Last, F.; Stone, O. Viruses in fungi and interferon stimulation. Nature 1968, 218, 542–545. [Google Scholar] [CrossRef]
  28. Van Alfen, N.; Jaynes, R.; Anagnostakis, S.; Day, P. Chestnut blight: Biological control by transmissible hypovirulence in Endothia parasitica. Science 1975, 189, 890–891. [Google Scholar] [CrossRef]
  29. Berry, E.; Bevan, E. A new species of double-stranded RNA from yeast. Nature 1972, 239, 279–280. [Google Scholar] [CrossRef]
  30. Maske, B.L.; Neto, D.P.D.C.; da Silva, G.B.; Lindner, J.D.D.; Soccol, C.R.; de Melo Pereira, G.V. Yeast viruses and their implications in fermented foods and beverages. Curr. Opin. Food Sci. 2022, 47, 100879. [Google Scholar] [CrossRef]
  31. Donaire, L.; Rozas, J.; Ayllón, M.A. Molecular characterization of Botrytis ourmia-like virus, a mycovirus close to the plant pathogenic genus Ourmiavirus. Virology 2016, 489, 158–164. [Google Scholar] [CrossRef][Green Version]
  32. Neupane, A.; Feng, C.; Feng, J.; Kafle, A.; Bücking, H.; Lee Marzano, S.-Y. Metatranscriptomic Analysis and In Silico Approach Identified Mycoviruses in the Arbuscular Mycorrhizal Fungus Rhizophagus spp. Viruses 2018, 10, 707. [Google Scholar] [CrossRef][Green Version]
  33. Nibert, M.L.; Ghabrial, S.A.; Maiss, E.; Lesker, T.; Vainio, E.J.; Jiang, D.; Suzuki, N. Taxonomic reorganization of family Partitiviridae and other recent progress in partitivirus research. Virus Res. 2014, 188, 128–141. [Google Scholar] [CrossRef]
  34. Li, C.X.; Zhu, J.Z.; Gao, B.D.; Zhu, H.J.; Zhou, Q.; Zhong, J. Characterization of a Novel Ourmia-Like Mycovirus Infecting Magnaporthe oryzae and Implications for Viral Diversity and Evolution. Viruses 2019, 11, 223. [Google Scholar] [CrossRef][Green Version]
  35. Kwon, S.-J.; Lim, W.-S.; Park, S.-H.; Park, M.-R.; Kim, K.-H. Molecular characterization of a dsRNA mycovirus, Fusarium graminearum virus-DK21, which is phylogenetically related to hypoviruses but has a genome organization and gene expression strategy resembling those of plant potex-like viruses. Mol. Cell 2007, 23, 304. [Google Scholar] [CrossRef]
  36. Andika, I.B.; Wei, S.; Cao, C.; Salaipeth, L.; Kondo, H.; Sun, L. Phytopathogenic fungus hosts a plant virus: A naturally occurring cross-kingdom viral infection. Proc. Natl. Acad. Sci. USA 2017, 114, 12267–12272. [Google Scholar] [CrossRef][Green Version]
  37. Cao, X.; Liu, J.; Pang, J.; Kondo, H.; Chi, S.; Zhang, J.; Sun, L.; Andika, I.B. Common but Nonpersistent Acquisitions of Plant Viruses by Plant-Associated Fungi. Viruses 2022, 14, 2279. [Google Scholar] [CrossRef]
  38. Roossinck, M.J. Evolutionary and ecological links between plant and fungal viruses. New Phytol. 2019, 221, 86–92. [Google Scholar] [CrossRef][Green Version]
  39. Wang, Q.; Zou, Q.; Dai, Z.; Hong, N.; Wang, G.; Wang, L. Four Novel Mycoviruses from the Hypovirulent Botrytis cinerea SZ-2-3y Isolate from Paris polyphylla: Molecular Characterisation and Mitoviral Sequence Transboundary Entry into Plants. Viruses 2022, 14, 151. [Google Scholar] [CrossRef]
  40. Bian, R.; Andika, I.B.; Pang, T.; Lian, Z.; Wei, S.; Niu, E.; Wu, Y.; Kondo, H.; Liu, X.; Sun, L. Facilitative and synergistic interactions between fungal and plant viruses. Proc. Natl. Acad. Sci. USA 2020, 117, 3779–3788. [Google Scholar] [CrossRef]
  41. Kartali, T.; Zsindely, N.; Nyilasi, I.; Németh, O.; Sávai, G.N.; Kocsubé, S.; Lipinszki, Z.; Patai, R.; Spisák, K.; Nagy, G.; et al. Molecular Characterization of Novel Mycoviruses in Seven Umbelopsis Strains. Viruses 2022, 14, 2343. [Google Scholar] [CrossRef]
  42. Petrzik, K.; Koloniuk, I.; Sehadová, H.; Sarkisova, T. Chrysoviruses inhabited symbiotic fungi of lichens. Viruses 2019, 11, 1120. [Google Scholar] [CrossRef] [PubMed][Green Version]
  43. Wan, X.; Zhao, Y.; Zhang, Y.; Wei, C.; Du, H.; Zhang, H.; Chen, J.; Yang, L.; Zang, R.; Wen, C. Molecular characterization of a novel partitivirus isolated from the phytopathogenic fungus Aplosporella javeedii. Arch. Virol. 2021, 166, 1237–1240. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, H.; Liu, M.; Zhu, H.; Zhong, J.; Liao, X.; Zhou, Q. Molecular characterization of a novel mitovirus from the plant-pathogenic fungus Botryosphaeria dothidea. Arch. Virol. 2021, 166, 633–637. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, L.; Jiang, J.; Wang, Y.; Hong, N.; Zhang, F.; Xu, W.; Wang, G. Hypovirulence of the phytopathogenic fungus Botryosphaeria dothidea: Association with a coinfecting chrysovirus and a partitivirus. J. Virol. 2014, 88, 7517–7527. [Google Scholar] [CrossRef][Green Version]
  46. Zhai, L.; Xiang, J.; Zhang, M.; Fu, M.; Yang, Z.; Hong, N.; Wang, G. Characterization of a novel double-stranded RNA mycovirus conferring hypovirulence from the phytopathogenic fungus Botryosphaeria dothidea. Virology 2016, 493, 75–85. [Google Scholar] [CrossRef]
  47. Yang, M.; Wang, G.; Mu, W.; Guo, J.; Xi, J.; Hu, L.; Liang, K.; Xu, Q.; Hu, Z.; Song, J. A novel victorivirus isolated from the tobacco spot blight fungus Stagonosporopsis cucurbitacearum in China. Arch. Virol. 2022, 167, 2845–2850. [Google Scholar] [CrossRef]
  48. He, Y.; Zou, Q.; Li, S.; Zhu, H.; Hong, N.; Wang, G.; Wang, L. Molecular characterization of a new fusarivirus infecting Botryosphaeria dothidea, the causal agent of pear ring rot disease. Arch. Virol. 2022, 167, 1893–1897. [Google Scholar] [CrossRef]
  49. Khan, H.A.; Telengech, P.; Kondo, H.; Bhatti, M.F.; Suzuki, N. Mycovirus hunting revealed the presence of diverse viruses in a single isolate of the phytopathogenic fungus Diplodia seriata from Pakistan. Front. Cell. Infect. Microbiol. 2022, 12, 702. [Google Scholar] [CrossRef]
  50. De Wet, J.; Bihon, W.; Preisig, O.; Wingfield, B.D.; Wingfield, M.J. Characterization of a novel dsRNA element in the pine endophytic fungus Diplodia scrobiculata. Arch. Virol. 2011, 156, 1199–1208. [Google Scholar] [CrossRef][Green Version]
  51. Aday, A.; Lehtijarvi, A.; Doğmuş-Lehtijarvi, H. Frequency of virus in some Diplodia pinea and Gremmeniella abietina isolates originated from Turkey. J. Agric. Ext. Rural Dev. 2012, 4, 181–183. [Google Scholar] [CrossRef]
  52. Jo, Y.; Choi, H.; Chu, H.; Cho, W.K. Unveiling Mycoviromes Using Fungal Transcriptomes. Int. J. Mol. Sci. 2022, 23, 10926. [Google Scholar] [CrossRef]
  53. Hafez, E.E.; Aseel, D.G.; Mostafa, S. Two novel mycoviruses related to Geminivirus isolated from the soil-borne fungi Macrophomina phaseolina (Tassi) Goid. and Mucor racemosus Bull. Biotechnol. Biotechnol. Equip. 2013, 27, 4222–4226. [Google Scholar] [CrossRef][Green Version]
  54. Wang, J.; Xiao, Y.; Zhao, H.; Ni, Y.; Liu, X.; Zhao, X.; Wang, G.; Xiao, X.; Liu, H. A novel double-stranded RNA mycovirus that infects Macrophomina phaseolina. Arch. Virol. 2019, 164, 2411–2416. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, J.; Ni, Y.; Liu, X.; Zhao, H.; Xiao, Y.; Xiao, X.; Li, S.; Liu, H. Divergent RNA viruses in Macrophomina phaseolina exhibit potential as virocontrol agents. Virus Evol. 2021, 7, veaa095. [Google Scholar] [CrossRef] [PubMed]
  56. Marzano, S.-Y.L.; Nelson, B.D.; Ajayi-Oyetunde, O.; Bradley, C.A.; Hughes, T.J.; Hartman, G.L.; Eastburn, D.M.; Domier, L.L. Identification of diverse mycoviruses through metatranscriptomics characterization of the viromes of five major fungal plant pathogens. J. Virol. 2016, 90, 6846–6863. [Google Scholar] [CrossRef] [PubMed][Green Version]
  57. Marais, A.; Faure, C.; Comont, G.; Candresse, T.; Stempien, E.; Corio-Costet, M.-F. Characterization of the mycovirome of the phytopathogenic fungus, Neofusicoccum parvum. Viruses 2021, 13, 375. [Google Scholar] [CrossRef]
  58. Marais, A.; Nivault, A.; Faure, C.; Comont, G.; Theil, S.; Candresse, T.; Corio-Costet, M.-F. Molecular characterization of a novel fusarivirus infecting the plant-pathogenic fungus Neofusicoccum luteum. Arch. Virol. 2018, 163, 559–562. [Google Scholar] [CrossRef]
  59. Marais, A.; Nivault, A.; Faure, C.; Theil, S.; Comont, G.; Candresse, T.; Corio-Costet, M.-F. Determination of the complete genomic sequence of Neofusicoccum luteum mitovirus 1 (NLMV1), a novel mitovirus associated with a phytopathogenic Botryosphaeriaceae. Arch. Virol. 2017, 162, 2477–2480. [Google Scholar] [CrossRef][Green Version]
  60. Zhao, Y.; Du, H.; Liu, Y.; Zhong, R.; Guan, Z.; Wang, G.; Zhang, Y.; Wei, C.; Wang, M.; Wan, X. Molecular characterization of a novel victorivirus isolated from the phytopathogenic fungus Phaeobotryon rhois. Arch. Virol. 2023, 168, 15. [Google Scholar] [CrossRef]
  61. Gilbert, K.B.; Holcomb, E.E.; Allscheid, R.L.; Carrington, J.C. Hiding in plain sight: New virus genomes discovered via a systematic analysis of fungal public transcriptomes. PLoS ONE 2019, 14, e0219207. [Google Scholar] [CrossRef][Green Version]
  62. McHale, M.T.; Roberts, I.N.; Noble, S.M.; Beaumont, C.; Whitehead, M.P.; Seth, D.; Oliver, R.P. CfT-I: An LTR-retrotransposon in Cladosporium fulvum, a fungal pathogen of tomato. Mol. Gener. Genet. 1992, 233, 337–347. [Google Scholar] [CrossRef]
  63. Nerva, L.; Turina, M.; Zanzotto, A.; Gardiman, M.; Gaiotti, F.; Gambino, G.; Chitarra, W. Isolation, molecular characterization and virome analysis of culturable wood fungal endophytes in esca symptomatic and asymptomatic grapevine plants. Environ. Microbiol. 2019, 21, 2886–2904. [Google Scholar] [CrossRef]
  64. Li, Y.; Zhou, M.; Yang, Y.; Liu, Q.; Zhang, Z.; Han, C.; Wang, Y. Characterization of the mycovirome from the plant-pathogenic fungus Cercospora beticola. Viruses 2021, 13, 1915. [Google Scholar] [CrossRef]
  65. Daudu, J.; Snowden, J.; Tubby, K.; Coutts, R.; Kotta-Loizou, I. Studying a mycovirus from Dothistroma septosporum, causative agent of pine needle blight. Access Microbiol 2019, 1, 783. [Google Scholar] [CrossRef]
  66. Lima, S.S.; Abadio, A.K.R.; Araujo, E.F.; Kitajima, E.W.; Sartorato, A.; Vieira de Queiroz, M. Mycovirus in Pseudocercospora griseola, the causal agent of angular leaf spot in common bean. Can. J. Microbiol. 2010, 56, 359–365. [Google Scholar] [CrossRef]
  67. Zelikovitch, N.; Eyal, Z.; Ben-Zvi, B.; Koltin, Y. Double-stranded RNA mycoviruses in Septoria tritici. Mycol. Res. 1990, 94, 590–594. [Google Scholar] [CrossRef]
  68. Cheng, J.; Jiang, D.; Fu, Y.; Li, G.; Peng, Y.; Ghabrial, S.A. Molecular characterization of a dsRNA totivirus infecting the sclerotial parasite Coniothyrium minitans. Virus Res. 2003, 93, 41–50. [Google Scholar] [CrossRef]
  69. Wang, Y.; Zhao, H.; Xue, C.; Xu, C.; Geng, Y.; Zang, R.; Guo, Y.; Wu, H.; Zhang, M. Complete genome sequence of a novel mycovirus isolated from the phytopathogenic fungus Corynespora cassiicola in China. Arch. Virol. 2020, 165, 2401–2404. [Google Scholar] [CrossRef]
  70. Petrzik, K.; Koloniuk, I.; Sarkisova, T.; Hrabáková, L. Detection and genome sequence of a new betapartitivirus associated with Cucurbitaria piceae Borthw. fungus causing bud blight of spruce in the Czech Republic. Arch. Virol. 2016, 161, 1405–1409. [Google Scholar] [CrossRef]
  71. Ye, L.; Shi, X.; He, Y.; Chen, J.; Xu, Q.; Shafik, K.; Xu, W. Characterization of a novel botybirnavirus with a unique dsRNA infecting Didymella theifolia from tea plants. bioRxiv 2022. [Google Scholar] [CrossRef]
  72. Zhou, J.; Hu, X.; Liang, X.; Wang, Y.; Xie, C.; Zheng, L. Complete genome sequence of a novel mycovirus from Phoma matteucciicola. Arch. Virol. 2021, 166, 317–320. [Google Scholar] [CrossRef] [PubMed]
  73. Zheng, F.; Xu, G.; Zhou, J.; Xie, C.; Cui, H.; Miao, W.; Kang, Z.; Zheng, L. Complete genomic sequence and organization of a novel mycovirus from Phoma matteuccicola strain LG915. Arch. Virol. 2019, 164, 2209–2213. [Google Scholar] [CrossRef] [PubMed]
  74. Zhou, J.; Wang, Y.; Liang, X.; Xie, C.; Liu, W.; Miao, W.; Kang, Z.; Zheng, L. Molecular characterization of a novel ourmia-like virus infecting Phoma matteucciicola. Viruses 2020, 12, 231. [Google Scholar] [CrossRef] [PubMed][Green Version]
  75. Zhou, S.; Chen, D.; Fu, Y.; Zhou, J.; Yang, Y.; Xie, C.; Zheng, L. Characterization of a novel mycotombus-like virus from the plant pathogenic fungus Phoma matteucciicola. Arch. Virol. 2023, 168, 103. [Google Scholar] [CrossRef] [PubMed]
  76. Shah, U.A.; Kotta-Loizou, I.; Fitt, B.D.; Coutts, R.H. Identification, molecular characterization, and biology of a novel quadrivirus infecting the phytopathogenic fungus Leptosphaeria biglobosa. Viruses 2018, 11, 9. [Google Scholar] [CrossRef][Green Version]
  77. Shah, U.A.; Kotta-Loizou, I.; Fitt, B.D.; Coutts, R.H. Mycovirus-induced hypervirulence of Leptosphaeria biglobosa enhances systemic acquired resistance to Leptosphaeria maculans in Brassica napus. Mol. Plant Microbe Interact. 2020, 33, 98–107. [Google Scholar] [CrossRef][Green Version]
  78. Wang, Y.; Xu, Z.; Cheng, J.; Fu, Y.; Jiang, D.; Xie, J. Characterization of a novel RNA virus from the phytopathogenic fungus Leptosphaeria biglobosa related to members of the genus Mitovirus. Arch. Virol. 2019, 164, 913–916. [Google Scholar] [CrossRef]
  79. Xie, J.; Havens, W.M.; Lin, Y.-H.; Suzuki, N.; Ghabrial, S.A. The victorivirus Helminthosporium victoriae virus 190S is the primary cause of disease/hypovirulence in its natural host and a heterologous host. Virus Res. 2016, 213, 238–245. [Google Scholar] [CrossRef]
  80. Dunkle, L.D. Double-stranded RNA mycovirus in Periconia circinata. Physiol. Plant Pathol. 1974, 4, 107–116. [Google Scholar] [CrossRef]
  81. Lin, Y.; Zhang, H.; Zhao, C.; Liu, S.; Guo, L. The complete genome sequence of a novel mycovirus from Alternaria longipes strain HN28. Arch. Virol. 2015, 160, 577–580. [Google Scholar] [CrossRef]
  82. Aoki, N.; Moriyama, H.; Kodama, M.; Arie, T.; Teraoka, T.; Fukuhara, T. A novel mycovirus associated with four double-stranded RNAs affects host fungal growth in Alternaria alternata. Virus Res. 2009, 140, 179–187. [Google Scholar] [CrossRef]
  83. Da Silva Xavier, A.; de Barros, A.P.O.; Godinho, M.T.; Zerbini, F.M.; de Oliveira Souza, F.; Bruckner, F.P.; Alfenas-Zerbini, P. A novel mycovirus associated to Alternaria alternata comprises a distinct lineage in Partitiviridae. Virus Res. 2018, 244, 21–26. [Google Scholar] [CrossRef]
  84. Okada, R.; Ichinose, S.; Takeshita, K.; Urayama, S.-i.; Fukuhara, T.; Komatsu, K.; Arie, T.; Ishihara, A.; Egusa, M.; Kodama, M. Molecular characterization of a novel mycovirus in Alternaria alternata manifesting two-sided effects: Down-regulation of host growth and up-regulation of host plant pathogenicity. Virology 2018, 519, 23–32. [Google Scholar] [CrossRef]
  85. Komatsu, K.; Katayama, Y.; Omatsu, T.; Mizutani, T.; Fukuhara, T.; Kodama, M.; Arie, T.; Teraoka, T.; Moriyama, H. Genome sequence of a novel victorivirus identified in the phytopathogenic fungus Alternaria arborescens. Arch. Virol. 2016, 161, 1701–1704. [Google Scholar] [CrossRef]
  86. Wang, W.; Wang, X.; Tu, C.; Yang, M.; Xiang, J.; Wang, L.; Hong, N.; Zhai, L.; Wang, G. Novel Mycoviruses Discovered from a Metatranscriptomics Survey of the Phytopathogenic Alternaria Fungus. Viruses 2022, 14, 2552. [Google Scholar] [CrossRef]
  87. Chen, Y.; Shang, H.H.; Yang, H.Q.; Da Gao, B.; Zhong, J. A mitovirus isolated from the phytopathogenic fungus Alternaria brassicicola. Arch. Virol. 2017, 162, 2869–2874. [Google Scholar] [CrossRef]
  88. Li, H.; Bian, R.; Liu, Q.; Yang, L.; Pang, T.; Salaipeth, L.; Andika, I.B.; Kondo, H.; Sun, L. Identification of a novel hypovirulence-inducing hypovirus from Alternaria alternata. Front. Microbiol. 2019, 10, 1076. [Google Scholar] [CrossRef][Green Version]
  89. Wang, H.; Li, C.; Cai, L.; Fang, S.; Zheng, L.; Yan, F.; Zhang, S.; Liu, Y. The complete genomic sequence of a novel botybirnavirus isolated from a phytopathogenic Bipolaris maydis. Virus Genes 2018, 54, 733–736. [Google Scholar] [CrossRef]
  90. Deng, Q.; Wang, H.; Li, C.; Li, P.; Fang, S.; Yang, S.; Yan, F.; Zhang, S.; Chen, Z. The complete genomic sequence of a novel alphapartitivirus from Bipolaris maydis, the causal agent of corn southern leaf blight. Arch. Virol. 2017, 162, 2433–2436. [Google Scholar] [CrossRef]
  91. Suikphopetch, P.; Suwanmanee, S.; Pumeesat, P.; Ampawong, S. In vitro characterization of Chrysovirus-1-induced hypovirulence of Bipolaris maydis. Walailak. J. Sci. Technol. 2021, 18, 6564. [Google Scholar] [CrossRef]
  92. Li, Q.; Huang, W.; Hai, D.; Wang, Y.; Xie, J.; Wang, M. The complete genome sequence of a novel hypovirus infecting Bipolaris oryzae. Arch. Virol. 2020, 165, 1027–1031. [Google Scholar] [CrossRef] [PubMed]
  93. Ghabrial, S.A.; Soldevila, A.I.; Havens, W.M. Molecular genetics of the viruses infecting the plant pathogenic fungus Helminthosporium victoriae. In dsRNA Genetic Elements: Concepts and Applications in Agriculture, Forestry, and Medicine; Tavantzis, S.M., Ed.; CRC Press: Boca Raton, FL, USA, 2002; pp. 213–236. [Google Scholar]
  94. Márquez, L.M.; Redman, R.S.; Rodriguez, R.J.; Roossinck, M.J. A virus in a fungus in a plant: Three-way symbiosis required for thermal tolerance. Science 2007, 315, 513–515. [Google Scholar] [CrossRef] [PubMed][Green Version]
  95. Shepherd, H.S. Characterization and localization of a virus-like particle in a Drechslera species. Exp. Mycol. 1990, 14, 294–298. [Google Scholar] [CrossRef]
  96. Livore, A.B.; Grubb, P.; Magill, C.W.; Magill, J.M. A Sephadex column procedure for DNA isolation is also useful for detecting dsRNA. Nucleic Acids Res. 1988, 16, 776. [Google Scholar] [CrossRef] [PubMed][Green Version]
  97. Zheng, Y.; Yin, S.; Zhao, Y.; Li, S.; Lu, Z.; Li, Z.; Deng, Q.; Li, Z.; Zhang, S.; Fang, S. Molecular and biological characteristics of a novel chrysovirus infecting the fungus phytopathogenic Setosphaeria turcica f.sp. sorghi. Virus Res. 2023, 325, 199037. [Google Scholar] [CrossRef]
  98. Nerva, L.; Ciuffo, M.; Vallino, M.; Margaria, P.; Varese, G.; Gnavi, G.; Turina, M. Multiple approaches for the detection and characterization of viral and plasmid symbionts from a collection of marine fungi. Virus Res. 2016, 219, 22–38. [Google Scholar] [CrossRef]
  99. Gao, Z.; Cai, L.; Liu, M.; Wang, X.; Yang, J.; An, H.; Deng, Q.; Zhang, S.; Fang, S. A novel previously undescribed fusarivirus from the phytopathogenic fungus Setosphaeria turcica. Arch. Virol. 2021, 166, 665–669. [Google Scholar] [CrossRef]
  100. Gao, Z.; Zhang, M.; Yu, T.; Wang, X.; Wang, X.; An, H.; Zhang, S.; Liu, M.; Fang, S. Molecular characterization of a novel polymycovirus from the phytopathogenic fungus Setosphaeria turcica. Arch. Virol. 2021, 166, 2315–2319. [Google Scholar] [CrossRef]
  101. Liu, H.; Wang, H.; Liao, X.L.; Gao, B.; Lu, X.; Sun, D.; Gong, W.; Zhong, J.; Zhu, H.; Pan, X. Mycoviral gene integration converts a plant pathogenic fungus into a biocontrol agent. Proc. Natl. Acad. Sci. USA 2022, 119, e2214096119. [Google Scholar] [CrossRef]
  102. McGinty, R.; Buck, K.; Rawlinson, C. Transcriptase activity assoclated with a type 2 double-stranded RNA mycovirus. Biochem. Biophys. Res. Commun. 1981, 98, 501–506. [Google Scholar] [CrossRef]
  103. Buck, K.; McGinty, R.; Rawlinson, C. Two Serologically Unrelated Viruses Isolated from a Phialophora sp. J. Gen. Virol. 1981, 55, 235–239. [Google Scholar] [CrossRef]
  104. Kanhayuwa, L.; Kotta-Loizou, I.; Özkan, S.; Gunning, A.P.; Coutts, R.H.A. A novel mycovirus from Aspergillus fumigatus contains four unique dsRNAs as its genome and is infectious as dsRNA. Proc. Natl. Acad. Sci. USA 2015, 112, 9100–9105. [Google Scholar] [CrossRef][Green Version]
  105. Kotta-Loizou, I.; Coutts, R.H. Mycoviruses in Aspergilli: A comprehensive review. Front. Microbiol. 2017, 8, 1699. [Google Scholar] [CrossRef][Green Version]
  106. Zoll, J.; Verweij, P.E.; Melchers, W.J. Discovery and characterization of novel Aspergillus fumigatus mycoviruses. PLoS ONE 2018, 13, e0200511. [Google Scholar] [CrossRef]
  107. Kinsella, C.M.; Deijs, M.; Gittelbauer, H.; van der Hoek, L.; van Dijk, K. Human clinical isolates of pathogenic fungi are host to diverse mycoviruses. Microbiol. Spectr. 2022, 10, e01610-22. [Google Scholar] [CrossRef]
  108. Niu, Y.; Yuan, Y.; Mao, J.; Yang, Z.; Cao, Q.; Zhang, T.; Wang, S.; Liu, D. Characterization of two novel mycoviruses from Penicillium digitatum and the related fungicide resistance analysis. Sci. Rep. 2018, 8, 5513. [Google Scholar] [CrossRef][Green Version]
  109. DeMarini, D.; Kurtzman, C.; Fennell, D.; Worden, K.; Detroy, R. Transmission of PsV-F and PsV-S mycoviruses during conidiogenesis of Penicillium stoloniferum. Microbiology 1977, 100, 59–64. [Google Scholar] [CrossRef][Green Version]
  110. Lau, S.K.; Lo, G.C.; Chow, F.W.; Fan, R.Y.; Cai, J.J.; Yuen, K.-Y.; Woo, P.C. Novel partitivirus enhances virulence of and causes aberrant gene expression in Talaromyces marneffei. MBio 2018, 9, e00947-18. [Google Scholar] [CrossRef][Green Version]
  111. Jiang, Y.; Tian, X.; Liu, X.; Yang, B.; Wang, N.; Wang, Q.; Yu, W.; Qi, X.; Peng, J.; Hsiang, T. Complete genome sequence of a novel chrysovirus infecting Talaromyces neofusisporus. Arch. Virol. 2022, 167, 2789–2793. [Google Scholar] [CrossRef]
  112. Teng, L.; Chen, S.; Hu, Z.; Chen, J.; Liu, H.; Zhang, T. Molecular characterization and transcriptomic analysis of a novel polymycovirus in the fungus Talaromyces amestolkiae. Front. Microbiol. 2022, 13, 4249. [Google Scholar] [CrossRef]
  113. Kohno, S.; Fujimura, T.; Rulong, S.; Kwon-Chung, K. Double-stranded RNA virus in the human pathogenic fungus Blastomyces dermatitidis. J. Virol. 1994, 68, 7554–7558. [Google Scholar] [CrossRef] [PubMed][Green Version]
  114. Li, Z.; Su, S.; Hamilton, M.; Yan, L.; Chen, Y. The ability to cause infection in a pathogenic fungus uncovers a new biological feature of honey bee viruses. J. Invertebr. Pathol. 2014, 120, 18–22. [Google Scholar] [CrossRef] [PubMed]
  115. Cheng, X.; Zhang, L.; Luo, J.; Yang, S.; Deng, Y.; Li, J.; Hou, C. Two pathogenic fungi isolated from chalkbrood samples and honey bee viruses they carried. Front. Microbiol. 2022, 13, 806. [Google Scholar] [CrossRef] [PubMed]
  116. Nerva, L.; Forgia, M.; Ciuffo, M.; Chitarra, W.; Chiapello, M.; Vallino, M.; Varese, G.; Turina, M. The mycovirome of a fungal collection from the sea cucumber Holothuria polii. Virus Res. 2019, 273, 197737. [Google Scholar] [CrossRef]
  117. Kim, J.-M.; Jung, J.-E.; Park, J.-A.; Park, S.-M.; Cha, B.-J.; Kim, D.-H. Biological function of a novel chrysovirus, CnV1-BS122, in the Korean Cryphonectria nitschkei BS122 strain. J. Biosci. Bioeng. 2013, 115, 1–3. [Google Scholar] [CrossRef]
  118. Deng, F.; Allen, T.D.; Hillman, B.I.; Nuss, D.L. Comparative Analysis of Alterations in Host Phenotype and Transcript Accumulation following Hypovirus and Mycoreovirus Infections of the Chestnut Blight Fungus Cryphonectria parasitica. Eukaryot. Cell. 2007, 6, 1286–1298. [Google Scholar] [CrossRef][Green Version]
  119. Cornejo, C.; Hisano, S.; Bragança, H.; Suzuki, N.; Rigling, D. A new double-stranded RNA mycovirus in Cryphonectria naterciae is able to cross the species barrier and is deleterious to a new host. J. Fungi 2021, 7, 861. [Google Scholar] [CrossRef]
  120. Nuskern, L.; Stojanović, M.; Milanović-Litre, M.; Šibenik, T.; Ježić, M.; Poljak, I.; Ćurković-Perica, M. Filling the Gap in Southern Europe—Diversity of Cryphonectria parasitica and Associated Mycovirus (Cryphonectria hypovirus 1) in Montenegro. J. Fungi 2022, 8, 552. [Google Scholar] [CrossRef]
  121. Rumbou, A.; Vainio, E.J.; Büttner, C. Towards the forest virome: High-throughput sequencing drastically expands our understanding on virosphere in temperate forest ecosystems. Microorganisms 2021, 9, 1730. [Google Scholar] [CrossRef]
  122. Hansen, D.R.; Van Alfen, N.K.; Gillies, K.; Powell, W.A. Naked dsRNA associated with hypovirulence of Endothia parasitica is packaged in fungal vesicles. J. Gen. Virol. 1985, 66, 2605–2614. [Google Scholar] [CrossRef]
  123. Preisig, O.; Moleleki, N.; Smit, W.; Wingfield, B.; Wingfield, M. A novel RNA mycovirus in a hypovirulent isolate of the plant pathogen Diaporthe ambigua. J. Gen. Virol. 2000, 81, 3107–3114. [Google Scholar] [CrossRef]
  124. Xu, G.; Zhang, X.; Liang, X.; Chen, D.; Xie, C.; Kang, Z.; Zheng, L. A novel hexa-segmented dsRNA mycovirus confers hypovirulence in the phytopathogenic fungus Diaporthe pseudophoenicicola. Environ. Microbiol. 2022, 24, 4274–4284. [Google Scholar] [CrossRef]
  125. Shafik, K.; Umer, M.; You, H.; Aboushedida, H.; Wang, Z.; Ni, D.; Xu, W. Characterization of a novel mitovirus infecting Melanconiella theae isolated from tea plants. Front. Microbiol. 2021, 12, 757556. [Google Scholar] [CrossRef]
  126. Peyambari, M.; Habibi, M.K.; Fotouhifar, K.-B.; Dizadji, A.; Roossinck, M.J. Molecular characterization of a novel putative partitivirus infecting Cytospora sacchari, a plant pathogenic fungus. Plant Pathol. J. 2014, 30, 151. [Google Scholar] [CrossRef][Green Version]
  127. Hrabáková, L.; Koloniuk, I.; Petrzik, K. Phomopsis longicolla RNA virus 1–Novel virus at the edge of myco-and plant viruses. Virology 2017, 506, 14–18. [Google Scholar] [CrossRef]
  128. Zhang, C.J.; Zhou, X.Y.; Zhong, J.; Guo, J.; Yang, X.P.; Zhu, H.J. Complete nucleotide sequence of a novel partitivirus infecting the plant-pathogenic fungus Phomopsis vexans. Arch. Virol. 2021, 166, 291–294. [Google Scholar] [CrossRef]
  129. Zhang, R.J.; Zhong, J.; Shang, H.H.; Pan, X.T.; Zhu, H.J.; Da Gao, B. The complete nucleotide sequence and genomic organization of a novel victorivirus with two non-overlapping ORFs, identified in the plant-pathogenic fungus Phomopsis vexans. Arch. Virol. 2015, 160, 1805–1809. [Google Scholar] [CrossRef]
  130. Xie, F.L.; Zhou, X.Y.; Xiao, R.; Zhang, C.J.; Zhong, J.; Zhou, Q.; Liu, F.; Zhu, H.J. Discovery and exploration of widespread infection of mycoviruses in Phomopsis vexans, the causal agent of phomopsis blight of eggplant in China. Front. Plant Sci. 2022, 13, 996862. [Google Scholar] [CrossRef]
  131. Yaegashi, H.; Kanematsu, S.; Ito, T. Molecular characterization of a new hypovirus infecting a phytopathogenic fungus, Valsa ceratosperma. Virus Res. 2012, 165, 143–150. [Google Scholar] [CrossRef]
  132. Xu, X.; Hai, D.; Li, J.; Huang, F.; Wang, Y. Molecular characterization of a novel penoulivirus from the phytopathogenic fungus Colletotrichum camelliae. Arch. Virol. 2022, 167, 641–644. [Google Scholar] [CrossRef]
  133. Guo, J.; Zhu, J.Z.; Zhou, X.Y.; Zhong, J.; Li, C.H.; Zhang, Z.G.; Zhu, H.J. A novel ourmia-like mycovirus isolated from the plant pathogenic fungus Colletotrichum gloeosporioides. Arch. Virol. 2019, 164, 2631–2635. [Google Scholar] [CrossRef] [PubMed]
  134. Suharto, A.R.; Jirakkakul, J.; Eusebio-Cope, A.; Salaipeth, L. Hypovirulence of Colletotrichum gloesporioides Associated with dsRNA Mycovirus Isolated from a Mango Orchard in Thailand. Viruses 2022, 14, 1921. [Google Scholar] [CrossRef] [PubMed]
  135. Fu, M.; Zhang, H.; Yin, M.; Han, Z.; Bai, Q.; Peng, Y.; Shafik, K.; Zhai, L.; Hong, N.; Xu, W. A novel heptasegmented positive-sense single-stranded RNA virus from the phytopathogenic fungus Colletotrichum fructicola. J. Virol. 2022, 96, e00318-22. [Google Scholar] [CrossRef] [PubMed]
  136. Zhu, J.Z.; Guo, J.; Hu, Z.; Zhang, X.T.; Li, X.G.; Zhong, J. A novel partitivirus that confer hypovirulence to the plant pathogenic fungus Colletotrichum liriopes. Front. Microbiol. 2021, 12, 653809. [Google Scholar] [CrossRef] [PubMed]
  137. Zhong, J.; Chen, D.; Lei, X.H.; Zhu, H.J.; Zhu, J.Z.; Da Gao, B. Detection and characterization of a novel Gammapartitivirus in the phytopathogenic fungus Colletotrichum acutatum strain HNZJ001. Virus Res. 2014, 190, 104–109. [Google Scholar] [CrossRef]
  138. Li, C.; Li, M.; Gao, Z.; Gong, D.; Hong, X.; Jiang, Y.; Chang, S.; Hu, M. Mycoviruses of Colletotrichum spp.: A review. South. J. Agric. 2020, 51, 123–132. [Google Scholar]
  139. Hrabáková, L.; Grum-Grzhimaylo, A.A.; Koloniuk, I.; Debets, A.J.; Sarkisova, T.; Petrzik, K. The alkalophilic fungus Sodiomyces alkalinus hosts beta-and gammapartitiviruses together with a new fusarivirus. PLoS ONE 2017, 12, e0187799. [Google Scholar] [CrossRef][Green Version]
  140. Cañizares, M.C.; López-Escudero, F.J.; Pérez-Artés, E.; García-Pedrajas, M.D. Characterization of a novel single-stranded RNA mycovirus related to invertebrate viruses from the plant pathogen Verticillium dahliae. Arch. Virol. 2018, 163, 771–776. [Google Scholar] [CrossRef]
  141. Feng, Z.; Zhu, H.; Li, Z.; Shi, Y.; Zhao, L.; Liu, L.; Jiang, D. Complete genome sequence of a novel dsRNA mycovirus isolated from the phytopathogenic fungus Verticillium dahliae Kleb. Arch. Virol. 2013, 158, 2621–2623. [Google Scholar] [CrossRef]
  142. Cañizares, M.C.; Pérez-Artés, E.; García-Pedrajas, M.D. The complete nucleotide sequence of a novel partitivirus isolated from the plant pathogenic fungus Verticillium albo-atrum. Arch. Virol. 2014, 159, 3141–3144. [Google Scholar] [CrossRef][Green Version]
  143. Gao, J.; Wang, N.; Jia, J.; Xie, J.; Jiang, D.; Zhang, L. Characterization of a novel magoulivirus isolated from the phytopathogenic fungus Verticillium dahlia. Arch. Virol. 2022, 167, 2387–2390. [Google Scholar] [CrossRef]
  144. Khalifa, M.E.; MacDiarmid, R.M. A novel totivirus naturally occurring in two different fungal genera. Front. Microbiol. 2019, 10, 2318. [Google Scholar] [CrossRef]
  145. Oh, C.-S.; Hillman, B.I. Genome organization of a partitivirus from the filamentous ascomycete Atkinsonella hypoxylon. J. Gen. Virol. 1995, 76, 1461–1470. [Google Scholar] [CrossRef]
  146. Romo, M.; Leuchtmann, A.; García, B.; Zabalgogeazcoa, I. Double-stranded RNA viruses infecting Epichloe festucae. N. Z. J. Agric. Res. 2006, 13, 207–209. [Google Scholar] [CrossRef]
  147. Romo, M.; Leuchtmann, A.; García, B.; Zabalgogeazcoa, I. A totivirus infecting the mutualistic fungal endophyte Epichloë festucae. Virus Res. 2007, 124, 38–43. [Google Scholar] [CrossRef]
  148. De la Paz Giménez-Pecci, M.; Bogo, M.R.; Santi, L.; Moraes, C.K.D.; Corrêa, C.T.; Henning Vainstein, M.; Schrank, A. Characterization of mycoviruses and analyses of chitinase secretion in the biocontrol fungus Metarhizium anisopliae. Curr. Microbiol. 2002, 45, 334–339. [Google Scholar] [CrossRef]
  149. Wang, P.; Yang, G.; Shi, N.; Huang, B. Molecular characterization of a novel double-stranded RNA virus infecting the entomopathogenic fungus Metarhizium brunneum. Arch. Microbiol. 2022, 204, 606. [Google Scholar] [CrossRef]
  150. Wang, P.; Yang, G.; Shi, N.; Huang, B. Molecular characterization of a new partitivirus, MbPV1, isolated from the entomopathogenic fungus Metarhizium brunneum in China. Arch. Virol. 2020, 165, 765–769. [Google Scholar] [CrossRef]
  151. He, L.; Wang, P.; Yang, G.; Chen, X.; Huang, B. A novel polymycovirus infecting the entomopathogenic fungus Metarhizium brunneum. Arch. Virol. 2023, 168, 6. [Google Scholar] [CrossRef]
  152. Jiang, Y.; Zhang, T.; Luo, C.; Jiang, D.; Li, G.; Li, Q.; Hsiang, T.; Huang, J. Prevalence and diversity of mycoviruses infecting the plant pathogen Ustilaginoidea virens. Virus Res. 2015, 195, 47–56. [Google Scholar] [CrossRef]
  153. Zhong, J.; Lei, X.H.; Zhu, J.Z.; Song, G.; Zhang, Y.D.; Chen, Y.; Gao, B.D. Detection and sequence analysis of two novel co-infecting double-strand RNA mycoviruses in Ustilaginoidea virens. Arch. Virol. 2014, 159, 3063–3070. [Google Scholar] [CrossRef] [PubMed]
  154. Zhu, H.J.; Chen, D.; Zhong, J.; Zhang, S.Y.; Gao, B.D. A novel mycovirus identified from the rice false smut fungus Ustilaginoidea virens. Virus Genes 2015, 51, 159–162. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, T.; Jiang, Y.; Dong, W. A novel monopartite dsRNA virus isolated from the phytopathogenic fungus Ustilaginoidea virens and ancestrally related to a mitochondria-associated dsRNA in the green alga Bryopsis. Virology 2014, 462, 227–235. [Google Scholar] [CrossRef] [PubMed][Green Version]
  156. Zhong, J.; Zhou, Q.; Lei, X.H.; Chen, D.; Shang, H.H.; Zhu, H.J. The nucleotide sequence and genome organization of two victoriviruses from the rice false smut fungus Ustilaginoidea virens. Virus Genes 2014, 48, 570–573. [Google Scholar] [CrossRef] [PubMed]
  157. Shi, N.; Hu, F.; Wang, P.; Zhang, Y.; Zhu, Q.; Yang, G.; Huang, B. Molecular characterization of two dsRNAs that could correspond to the genome of a new mycovirus that infects the entomopathogenic fungus Beauveria bassiana. Arch. Virol. 2021, 166, 3233–3237. [Google Scholar] [CrossRef]
  158. Filippou, C.; Diss, R.M.; Daudu, J.O.; Coutts, R.H.; Kotta-Loizou, I. The polymycovirus-mediated growth enhancement of the entomopathogenic fungus Beauveria bassiana is dependent on carbon and nitrogen metabolism. Front. Microbiol. 2021, 12, 606366. [Google Scholar] [CrossRef]
  159. Koloniuk, I.; Hrabáková, L.; Petrzik, K. Molecular characterization of a novel amalgavirus from the entomopathogenic fungus Beauveria bassiana. Arch. Virol. 2015, 160, 1585–1588. [Google Scholar] [CrossRef]
  160. Kotta-Loizou, I.; Coutts, R.H.A. Studies on the Virome of the Entomopathogenic Fungus Beauveria bassiana Reveal Novel dsRNA Elements and Mild Hypervirulence. PLoS Pathog. 2017, 13, e1006183. [Google Scholar] [CrossRef][Green Version]
  161. Zhang, Y.; Shi, N.; Wang, P.; Zhu, Q.; Yang, G.; Huang, B. Molecular characterization of a novel alternavirus infecting the entomopathogenic fungus Cordyceps chanhua. Arch. Virol. 2022, 167, 1467–1470. [Google Scholar] [CrossRef]
  162. Zhu, Q.; Shi, N.; Wang, P.; Zhang, Y.; Peng, F.; Yang, G.; Huang, B. A Novel Gammapartitivirus That Causes Changes in Fungal Development and Multi-Stress Tolerance to Important Medicinal Fungus Cordyceps chanhua. J. Fungi 2022, 8, 1309. [Google Scholar] [CrossRef]
  163. Herrero, N. Identification and sequence determination of a new chrysovirus infecting the entomopathogenic fungus Isaria javanica. Arch. Virol. 2017, 162, 1113–1117. [Google Scholar] [CrossRef]
  164. Petrzik, K.; Siddique, A.B. A mycoparasitic and opportunistic fungus is inhabited by a mycovirus. Arch. Virol. 2019, 164, 2545–2549. [Google Scholar] [CrossRef]
  165. Liu, C.; Li, M.; Redda, E.T.; Mei, J.; Zhang, J.; Wu, B.; Jiang, X. A novel double-stranded RNA mycovirus isolated from Trichoderma harzianum. Virol. J. 2019, 16, 113. [Google Scholar] [CrossRef]
  166. Chun, J.; Yang, H.-E.; Kim, D.-H. Identification and Molecular Characterization of a Novel Partitivirus from Trichoderma atroviride NFCF394. Viruses 2018, 10, 578. [Google Scholar] [CrossRef][Green Version]
  167. Lee, S.H.; Yun, S.-H.; Chun, J.; Kim, D.-H. Characterization of a novel dsRNA mycovirus of Trichoderma atroviride NFCF028. Arch. Virol. 2017, 162, 1073–1077. [Google Scholar] [CrossRef]
  168. Yun, S.-H.; Lee, S.H.; So, K.-K.; Kim, J.-M.; Kim, D.-H. Incidence of diverse dsRNA mycoviruses in Trichoderma spp. causing green mold disease of shiitake Lentinula edodes. FEMS Microbiol. Lett. 2016, 363, fnw220. [Google Scholar] [CrossRef][Green Version]
  169. Li, P.; Bhattacharjee, P.; Wang, S.; Zhang, L.; Ahmed, I.; Guo, L. Mycoviruses in Fusarium species: An update. Front. Cell. Infect. Microbiol. 2019, 9, 257. [Google Scholar] [CrossRef][Green Version]
  170. Zhao, Y.; Zhang, Y.; Wan, X.; She, Y.; Li, M.; Xi, H.; Xie, J.; Wen, C. A novel ourmia-like mycovirus confers hypovirulence-associated traits on Fusarium oxysporum. Front. Microbiol. 2020, 11, 569869. [Google Scholar] [CrossRef]
  171. Wang, J.; Li, C.; Song, P.; Qiu, R.; Song, R.; Li, X.; Ni, Y.; Zhao, H.; Liu, H.; Li, S. Molecular and Biological Characterization of the First Mymonavirus Identified in Fusarium oxysporum. Front. Microbiol. 2022, 13, 870204. [Google Scholar] [CrossRef]
  172. Ma, G.; Zhang, B.; Qi, K.; Zhang, Y.; Ma, L.; Jiang, H.; Qin, S.; Qi, J. Characterization of a novel mitovirus isolated from the phytopathogenic fungus Fusarium pseudograminearum. Arch. Virol. 2022, 167, 1369–1373. [Google Scholar] [CrossRef]
  173. Jacquat, A.G.; Theumer, M.G.; Cañizares, M.C.; Debat, H.J.; Iglesias, J.; García Pedrajas, M.D.; Dambolena, J.S. A survey of mycoviral infection in Fusarium spp. isolated from maize and sorghum in argentina identifies the first mycovirus from Fusarium verticillioides. Viruses 2020, 12, 1161. [Google Scholar] [CrossRef] [PubMed]
  174. Zhang, L.; Chen, X.; Bhattacharjee, P.; Shi, Y.; Guo, L.; Wang, S. Molecular characterization of a novel strain of Fusarium graminearum virus 1 infecting Fusarium graminearum. Viruses 2020, 12, 357. [Google Scholar] [CrossRef] [PubMed][Green Version]
  175. Li, P.; Wang, S.; Zhang, L.; Qiu, D.; Zhou, X.; Guo, L. A tripartite ssDNA mycovirus from a plant pathogenic fungus is infectious as cloned DNA and purified virions. Sci. Adv. 2020, 6, eaay9634. [Google Scholar] [CrossRef] [PubMed][Green Version]
  176. Lutz, T.; Langer, G.; Heinze, C. Complete genome sequence of a novel alternavirus infecting the fungus Ilyonectria crassa. Arch. Virol. 2023, 168, 34. [Google Scholar] [CrossRef]
  177. Pielhop, T.P.; Popp, C.; Knierim, D.; Margaria, P.; Maiß, E. First report of a chrysovirus infecting a member of the fungal genus Ilyonectria. Arch. Virol. 2022, 167, 2411–2415. [Google Scholar] [CrossRef]
  178. Ahn, I.-P.; Lee, Y.-H. A viral double-stranded RNA up regulates the fungal virulence of Nectria radicicola. Mol. Plant Microbe Interact. 2001, 14, 496–507. [Google Scholar] [CrossRef][Green Version]
  179. Pielhop, T.P.; Popp, C.; Knierim, D.; Margaria, P.; Mai, E. Three new mycoviruses identified in the apple replant disease (ARD)-associated fungus Rugonectria rugulosa. Virus Genes 2022, 58, 423–435. [Google Scholar] [CrossRef]
  180. Lutz, T.; Langer, G.; Heinze, C. Complete genome sequence of a new quadrivirus infecting a member of the genus Thelonectria. Arch. Virol. 2022, 167, 691–694. [Google Scholar] [CrossRef]
  181. Herrero, N.; Zabalgogeazcoa, I. Mycoviruses infecting the endophytic and entomopathogenic fungus Tolypocladium cylindrosporum. Virus Res. 2011, 160, 409–413. [Google Scholar] [CrossRef][Green Version]
  182. Herrero, N.; Sánchez Márquez, S.; Zabalgogeazcoa, I. Mycoviruses are common among different species of endophytic fungi of grasses. Arch. Virol. 2009, 154, 327–330. [Google Scholar] [CrossRef]
  183. Chen, H.; Jin, L.; Jiang, X.; Yu, Z.; Duns, G.J.; Shao, R.; Xu, W.; Chen, J. A novel mitovirus from Buergenerula spartinae infecting the invasive species Spartina alterniflora. Virol. Sin. 2014, 29, 257–260. [Google Scholar] [CrossRef]
  184. Jamil, N.; Buck, K.; Carlile, M. Sequence relationships between virus double-stranded RNA from isolates of Gaeumannomyces graminis in different vegetative compatibility groups. J. Gen. Virol. 1984, 65, 1741–1747. [Google Scholar] [CrossRef]
  185. Chun, S.J.; Lee, Y.-H. Inheritance of dsRNAs in the rice blast fungus, Magnaporthe grisea. FEMS Microbiol. Lett. 1997, 148, 159–162. [Google Scholar] [CrossRef]
  186. Aihara, M.; Urayama, S.-I.; Le, M.T.; Katoh, Y.; Higashiura, T.; Fukuhara, T.; Arie, T.; Teraoka, T.; Komatsu, K.; Moriyama, H. Infection by Magnaporthe oryzae chrysovirus 1 strain A triggers reduced virulence and pathogenic race conversion of its host fungus, Magnaporthe oryzae. J. Gen. Plant Pathol. 2018, 84, 92–103. [Google Scholar] [CrossRef]
  187. Lin, Y.; Zhou, J.; Zhou, X.; Shuai, S.; Zhou, R.; An, H.; Fang, S.; Zhang, S.; Deng, Q. A novel narnavirus from the plant-pathogenic fungus Magnaporthe oryzae. Arch. Virol. 2020, 165, 1235–1240. [Google Scholar] [CrossRef]
  188. Urayama, S.; Kato, S.; Suzuki, Y.; Aoki, N.; Le, M.T.; Arie, T.; Teraoka, T.; Fukuhara, T.; Moriyama, H. Mycoviruses related to chrysovirus affect vegetative growth in the rice blast fungus Magnaporthe oryzae. J. Gen. Virol. 2010, 91, 3085–3094. [Google Scholar] [CrossRef]
  189. Park, Y.; Chen, X.; Punja, Z.K. Molecular and biological characterization of a mitovirus in Chalara elegans (Thielaviopsis basicola). Phytopathology 2006, 96, 468–479. [Google Scholar] [CrossRef][Green Version]
  190. Park, Y.; James, D.; Punja, Z.K. Co-infection by two distinct totivirus-like double-stranded RNA elements in Chalara elegans (Thielaviopsis basicola). Virus Res. 2005, 109, 71–85. [Google Scholar] [CrossRef]
  191. Espinal, R.B.A.; de Santana, S.F.; Santos, V.C.; Lizardo, G.N.R.; Silva, R.J.S.; Corrêa, R.X.; Loguercio, L.L.; Neto, A.G.; Pirovani, C.P.; Fonseca, P.L.C. Uncovering a Complex Virome Associated with the Cacao Pathogens Ceratocystis cacaofunesta and Ceratocystis fimbriata. Pathogens 2023, 12, 287. [Google Scholar] [CrossRef]
  192. Deng, F.; Boland, G.J. Natural occurrence of a partitivirus in the sapstaining fungus Ceratocystis resinifera. Can. J. Plant Pathol. 2007, 29, 182–189. [Google Scholar] [CrossRef]
  193. Ohkita, S.; Lee, Y.; Nguyen, Q.; Ikeda, K.; Suzuki, N.; Nakayashiki, H. Three ourmia-like viruses and their associated RNAs in Pyricularia oryzae. Virology 2019, 534, 25–35. [Google Scholar] [CrossRef] [PubMed]
  194. Hunst, P.L.; Latterell, F.M.; Rossi, A.E. Variation in double-stranded RNA from isolates of Pyricularia oryzae. Phytopathology 1986, 76, 674–678. [Google Scholar] [CrossRef]
  195. Cole, T.E.; McLler, B.M.; Hong, Y.; Brasier, C.M.; Buck, K.W. Complexity of Virus-like Double-stranded RN A Elements in a Diseased Isolate of the Dutch Elm Disease Fungus, Ophiostoma novo-ulmi. J. Phytopathol. 1998, 146, 593–598. [Google Scholar] [CrossRef]
  196. Zhu, Y.; Lu, A.; Wang, Z.; Lu, Q.; Yu, C.; Liu, H. Molecular characterization of a novel endornavirus isolated from Ophiostoma bicolor associated with bark beetles. Arch. Virol. 2022, 167, 2839–2843. [Google Scholar] [CrossRef] [PubMed]
  197. Hong, Y.; Dover, S.L.; Cole, T.E.; Brasier, C.M.; Buck, K.W. Multiple Mitochondrial Viruses in an Isolate of the Dutch Elm Disease Fungus Ophiostoma Novo-Ulmi. Virology 1999, 258, 118–127. [Google Scholar] [CrossRef][Green Version]
  198. Honda, S.; Eusebio-Cope, A.; Miyashita, S.; Yokoyama, A.; Aulia, A.; Shahi, S.; Kondo, H.; Suzuki, N. Establishment of Neurospora crassa as a model organism for fungal virology. Nat. Commun. 2020, 11, 5627. [Google Scholar] [CrossRef]
  199. Zhou, Q.; Zhong, J.; Hu, Y.; Da Gao, B. A novel nonsegmented double-stranded RNA mycovirus identified in the phytopathogenic fungus Nigrospora oryzae shows similarity to partitivirus-like viruses. Arch. Virol. 2016, 161, 229–232. [Google Scholar] [CrossRef]
  200. Zhong, J.; Zhao, S.Q.; Li, G.F.; Pang, X.D.; Deng, X.J.; Zhu, H.J.; Da Gao, B.; Zhou, Q. A novel fusarivirus isolated from the phytopathogenic fungus Nigrospora oryzae. Virus Genes 2016, 52, 891–895. [Google Scholar] [CrossRef]
  201. Liu, H.; Liu, R.; Li, C.X.; Wang, H.; Zhu, H.J.; Gao, B.D.; Zhou, Q.; Zhong, J. A victorivirus and two novel mitoviruses co-infected the plant pathogen Nigrospora oryzae. Viruses 2019, 11, 83. [Google Scholar] [CrossRef][Green Version]
  202. Zhong, J.; Sui, W.W.; Li, P.; Tang, Q.J.; Liu, T.B.; Xiao, Y.S. Characterization of a novel victorivirus from Nigrospora chinensis, a fungus isolated from tobacco. Arch. Virol. 2022, 167, 2851–2855. [Google Scholar] [CrossRef]
  203. Armengol, J.; Alaniz, S.; Vicent, A.; Beltrán, R.; Abad-Campos, P.; Pérez-Sierra, A.; García-Jiménez, J.; Salem, I.B.; Souli, M.; Boughalleb, N. Effect of dsRNA on growth rate and reproductive potential of Monosporascus cannonballus. Fungal Biol. 2011, 115, 236–244. [Google Scholar] [CrossRef]
  204. Sahin, E.; Akata, I. Viruses infecting macrofungi. Virusdisease 2018, 29, 1–18. [Google Scholar] [CrossRef]
  205. De Rezende, R.R.; de Oliveira Souza, F.; Leal, L.L.; Morgan, T.; Bermudez, J.M.M.; Cascardo, R.d.S.; de Barros, D.R.; Alfenas-Zerbini, P. Characterization of a new mitovirus infecting the phytopathogenic fungus Microdochium albescens. Arch. Virol. 2021, 166, 3455–3459. [Google Scholar] [CrossRef]
  206. Chen, F.; Pu, Z.; Ni, H.; Wang, Y.; Yan, B. Multiple mycoviruses identified in Pestalotiopsis spp. from Chinese bayberry. Virol. J. 2021, 18, 43. [Google Scholar] [CrossRef]
  207. Zhou, L.; Li, X.; Kotta-Loizou, I.; Dong, K.; Li, S.; Ni, D.; Hong, N.; Wang, G.; Xu, W. A mycovirus modulates the endophytic and pathogenic traits of a plant associated fungus. ISME J. 2021, 15, 1893–1906. [Google Scholar] [CrossRef]
  208. Velasco, L.; Arjona-Girona, I.; Cretazzo, E.; López-Herrera, C. Viromes in Xylariaceae fungi infecting avocado in Spain. Virology 2019, 532, 11–21. [Google Scholar] [CrossRef]
  209. Velasco, L.; López-Herrera, C.; Cretazzo, E. Two novel partitiviruses that accumulate differentially in Rosellinia necatrix and Entoleuca sp. infecting avocado. Virus Res. 2020, 285, 198020. [Google Scholar] [CrossRef]
  210. Arjona-López, J.M.; López-Herrera, C.J. Entoleuca sp. infected by mycoviruses as potential biocontrol agents of avocado white root rot. Eur. J. Plant Pathol. 2021, 159, 409–420. [Google Scholar] [CrossRef]
  211. Chiba, S.; Salaipeth, L.; Lin, Y.-H.; Sasaki, A.; Kanematsu, S.; Suzuki, N. A novel bipartite double-stranded RNA mycovirus from the white root rot fungus Rosellinia necatrix: Molecular and biological characterization, taxonomic considerations, and potential for biological control. J. Virol. 2009, 83, 12801–12812. [Google Scholar] [CrossRef][Green Version]
  212. Lin, Y.-H.; Chiba, S.; Tani, A.; Kondo, H.; Sasaki, A.; Kanematsu, S.; Suzuki, N. A novel quadripartite dsRNA virus isolated from a phytopathogenic filamentous fungus, Rosellinia necatrix. Virology 2012, 426, 42–50. [Google Scholar] [CrossRef][Green Version]
  213. Arjona-López, J.M.; Telengech, P.; Suzuki, N.; López-Herrera, C.J. Coinfection of Rosellinia necatrix by a partitivirus and a virga-like virus is associated with hypovirulence. Eur. J. Plant Pathol. 2020, 158, 111–119. [Google Scholar] [CrossRef]
  214. Sasaki, A.; Nakamura, H.; Suzuki, N.; Kanematsu, S. Characterization of a new megabirnavirus that confers hypovirulence with the aid of a co-infecting partitivirus to the host fungus, Rosellinia necatrix. Virus Res. 2016, 219, 73–82. [Google Scholar] [CrossRef] [PubMed]
  215. Arjona-López, J.M.; Telengech, P.; Suzuki, N.; López-Herrera, C.J. A moderate level of hypovirulence conferred by a hypovirus in the avocado white root rot fungus, Rosellinia necatrix. Fungal Biol. 2021, 125, 69–76. [Google Scholar] [CrossRef] [PubMed]
  216. Kanematsu, S.; Arakawa, M.; Oikawa, Y.; Onoue, M.; Osaki, H.; Nakamura, H.; Ikeda, K.; Kuga-Uetake, Y.; Nitta, H.; Sasaki, A.; et al. A Reovirus Causes Hypovirulence of Rosellinia necatrix. Phytopathology 2004, 94, 561–568. [Google Scholar] [CrossRef][Green Version]
  217. Yaegashi, H.; Kanematsu, S. Natural infection of the soil-borne fungus Rosellinia necatrix with novel mycoviruses under greenhouse conditions. Virus Res. 2016, 219, 83–91. [Google Scholar] [CrossRef]
  218. Sutela, S.; Forgia, M.; Vainio, E.J.; Chiapello, M.; Daghino, S.; Vallino, M.; Martino, E.; Girlanda, M.; Perotto, S.; Turina, M. The virome from a collection of endomycorrhizal fungi reveals new viral taxa with unprecedented genome organization. Virus Evol. 2020, 6, veaa076. [Google Scholar] [CrossRef]
  219. Ren, P.; Rajkumar, S.S.; Zhang, T.; Sui, H.; Masters, P.S.; Martinkova, N.; Kubátová, A.; Pikula, J.; Chaturvedi, S.; Chaturvedi, V. A common partitivirus infection in United States and Czech Republic isolates of bat white-nose syndrome fungal pathogen Pseudogymnoascus destructans. Sci. Rep. 2020, 10, 13893. [Google Scholar] [CrossRef]
  220. Galán-Cubero, R.; Córdoba, L.; Rodríguez-Romero, J.; Chiapello, M.; Turina, M.; Ayllón, M.A. Molecular Data of a Novel Penoulivirus Associated with the Plant-Pathogenic Fungus Erysiphe necator. Phytopathology 2022, 112, 1587–1591. [Google Scholar] [CrossRef]
  221. Pandey, B.; Naidu, R.; Grove, G. Detection and analysis of mycovirus-related RNA viruses from grape powdery mildew fungus Erysiphe necator. Arch. Virol. 2018, 163, 1019–1030. [Google Scholar] [CrossRef]
  222. Botella, L.; Hantula, J. Description, distribution, and relevance of viruses of the forest pathogen Gremmeniella abietina. Viruses 2018, 10, 654. [Google Scholar] [CrossRef][Green Version]
  223. Botella, L.; Tuomivirta, T.T.; Vervuurt, S.; Diez, J.J.; Hantula, J. Occurrence of two different species of mitoviruses in the European race of Gremmeniella abietina var. abietina, both hosted by the genetically unique Spanish population. Fungal Biol. 2012, 116, 872–882. [Google Scholar] [CrossRef]
  224. Schoebel, C.N.; Zoller, S.; Rigling, D. Detection and genetic characterisation of a novel mycovirus in Hymenoscyphus fraxineus, the causal agent of ash dieback. Infect. Genet. Evol. 2014, 28, 78–86. [Google Scholar] [CrossRef]
  225. Shamsi, W.; Kondo, H.; Ulrich, S.; Rigling, D.; Prospero, S. Novel RNA viruses from the native range of Hymenoscyphus fraxineus, the causal fungal agent of ash dieback. Virus Res. 2022, 320, 198901. [Google Scholar] [CrossRef]
  226. Castro, M.; Kramer, K.; Valdivia, L.; Ortiz, S.; Castillo, A. A double-stranded RNA mycovirus confers hypovirulence-associated traits to Botrytis cinerea. FEMS Microbiol. Lett. 2003, 228, 87–91. [Google Scholar] [CrossRef]
  227. Howitt, R.L.J.; Beever, R.E.; Pearson, M.N.; Forster, R.L.S. Genome characterization of a flexuous rod-shaped mycovirus, Botrytis virus X, reveals high amino acid identity to genes from plant ‘potex-like’ viruses. Arch. Virol. 2006, 151, 563–579. [Google Scholar] [CrossRef]
  228. Pearson, M.N.; Bailey, A.M. Viruses of botrytis. In Advances in Virus Research; Academic Press: New York, NY, USA, 2013; Volume 86, pp. 249–272. [Google Scholar]
  229. Donaire, L.; Pagán, I.; Ayllón, M.A. Characterization of Botrytis cinerea negative-stranded RNA virus 1, a new mycovirus related to plant viruses, and a reconstruction of host pattern evolution in negative-sense ssRNA viruses. Virology 2016, 499, 212–218. [Google Scholar] [CrossRef]
  230. Hao, F.; Wu, M.; Li, G. Molecular characterization and geographic distribution of a mymonavirus in the population of Botrytis cinerea. Viruses 2018, 10, 432. [Google Scholar] [CrossRef][Green Version]
  231. Potgieter, C.A.; Castillo, A.; Castro, M.; Cottet, L.; Morales, A. A wild-type Botrytis cinerea strain co-infected by double-stranded RNA mycoviruses presents hypovirulence-associated traits. Virol. J. 2013, 10, 220. [Google Scholar] [CrossRef][Green Version]
  232. Pi-Fang, T.; Pearson, M.N.; Beever, R.E. Mycoviruses in Monilinia fructicola. Mycol. Res. 2004, 108, 907–912. [Google Scholar]
  233. De Miccolis Angelini, R.M.; Raguseo, C.; Rotolo, C.; Gerin, D.; Faretra, F.; Pollastro, S. The Mycovirome in a Worldwide Collection of the Brown Rot Fungus Monilinia fructicola. J. Fungi 2022, 8, 481. [Google Scholar] [CrossRef]
  234. Hamid, M.; Xie, J.; Wu, S.; Maria, S.; Zheng, D.; Assane Hamidou, A.; Wang, Q.; Cheng, J.; Fu, Y.; Jiang, D. A Novel Deltaflexivirus that Infects the Plant Fungal Pathogen, Sclerotinia sclerotiorum, Can Be Transmitted Among Host Vegetative Incompatible Strains. Viruses 2018, 10, 295. [Google Scholar] [CrossRef] [PubMed][Green Version]
  235. Li, K.; Zheng, D.; Cheng, J.; Chen, T.; Fu, Y.; Jiang, D.; Xie, J. Characterization of a novel Sclerotinia sclerotiorum RNA virus as the prototype of a new proposed family within the order Tymovirales. Virus Res. 2016, 219, 92–99. [Google Scholar] [CrossRef] [PubMed]
  236. Mu, F.; Jia, J.; Xue, Y.; Jiang, D.; Fu, Y.; Cheng, J.; Lin, Y.; Xie, J. Characterization of a novel botoulivirus isolated from the phytopathogenic fungus Sclerotinia sclerotiorum. Arch. Virol. 2021, 166, 2859–2863. [Google Scholar] [CrossRef] [PubMed]
  237. Xie, J.; Wei, D.; Jiang, D.; Fu, Y.; Li, G.; Ghabrial, S.; Peng, Y. Characterization of debilitation-associated mycovirus infecting the plant-pathogenic fungus Sclerotinia sclerotiorum. J. Gen. Virol. 2006, 87, 241–249. [Google Scholar] [CrossRef]
  238. Liu, L.; Xie, J.; Cheng, J.; Fu, Y.; Li, G.; Yi, X.; Jiang, D. Fungal negative-stranded RNA virus that is related to bornaviruses and nyaviruses. Proc. Natl. Acad. Sci. USA 2014, 111, 12205–12210. [Google Scholar] [CrossRef][Green Version]
  239. Zhong, J.; Chen, D.; Zhu, H.J.; Gao, B.D.; Zhou, Q. Hypovirulence of Sclerotium rolfsii caused by associated RNA mycovirus. Front. Microbiol. 2016, 7, 1798. [Google Scholar] [CrossRef][Green Version]
  240. Zhu, J.Z.; Zhu, H.J.; Gao, B.D.; Zhou, Q.; Zhong, J. Diverse, novel mycoviruses from the virome of a hypovirulent Sclerotium rolfsii strain. Front. Plant Sci. 2018, 9, 1738. [Google Scholar] [CrossRef][Green Version]
  241. Sahin, E.; Keskin, E.; Akata, I. Full-length genome characterization of a new fusagravirus hosted by the spring orange peel fungus Caloscypha fulgens. Acta Virol. 2021, 65, 313–319. [Google Scholar] [CrossRef]
  242. Sahin, E.; Keskin, E.; Akata, I. Molecular characterization of a novel partitivirus hosted by the false morel mushroom Gyromitra esculenta. Arch. Virol. 2021, 166, 1247–1251. [Google Scholar] [CrossRef]
  243. Sahin, E.; Keskin, E.; Akata, I. The unique genome organization of two novel fusariviruses hosted by the true morel mushroom Morchella esculenta. Virus Res. 2021, 302, 198486. [Google Scholar] [CrossRef]
  244. Sahin, E.; Akata, I.; Keskin, E. Novel and divergent bipartite mycoviruses associated with the ectomycorrhizal fungus Sarcosphaera coronaria. Virus Res. 2020, 286, 198071. [Google Scholar] [CrossRef]
  245. Sahin, E.; Ozbey Saridogan, B.G.; Keskin, E.; Akata, I. Identification and complete genome sequencing of a novel betapartitivirus naturally infecting the mycorrhizal desert truffle Terfezia claveryi. Virus Genes 2023, 59, 254–259. [Google Scholar] [CrossRef]
  246. Sahin, E.; Akata, I. Complete genome sequence of a novel mitovirus from the ectomycorrhizal fungus Geopora sumneriana. Arch. Virol. 2019, 164, 2853–2857. [Google Scholar] [CrossRef]
  247. Sahin, E.; Keskin, E.; Akata, I. Novel and diverse mycoviruses co-inhabiting the hypogeous ectomycorrhizal fungus Picoa juniperi. Virology 2021, 552, 10–19. [Google Scholar] [CrossRef]
  248. Stielow, B.; Klenk, H.-P.; Winter, S.; Menzel, W. A novel Tuber aestivum (Vittad.) mitovirus. Arch. Virol. 2011, 156, 1107–1110. [Google Scholar] [CrossRef]
  249. Stielow, B.; Menzel, W. Complete nucleotide sequence of TaV1, a novel totivirus isolated from a black truffle ascocarp (Tuber aestivum Vittad.). Arch. Virol. 2010, 155, 2075–2078. [Google Scholar] [CrossRef]
  250. Riccioni, C.; Rubini, A.; Belfiori, B.; Passeri, V.; Paolocci, F.; Arcioni, S. Tmt1: The first LTR-retrotransposon from a Tuber spp. Curr. Genet. 2008, 53, 23–34. [Google Scholar] [CrossRef]
  251. Taylor, D.J.; Ballinger, M.J.; Bowman, S.M.; Bruenn, J.A. Virus-host co-evolution under a modified nuclear genetic code. PeerJ 2013, 1, e50. [Google Scholar] [CrossRef][Green Version]
  252. Pospíšek, M.; Palkova, Z.; Korb, J.; Vaněk, D. Isolation and characterization of a new dsRNA virus from Wickerhamia fluorescens. Folia Microbiol. 1996, 41, 223–227. [Google Scholar] [CrossRef]
  253. Khan, H.A.; Kondo, H.; Shahi, S.; Bhatti, M.F.; Suzuki, N. Identification of novel totiviruses from the ascomycetous fungus Geotrichum candidum. Arch. Virol. 2022, 167, 2833–2838. [Google Scholar] [CrossRef]
  254. Matte, O.; Chabalier, C.; Ratomahenina, R.; Bossy, J.; Galzy, P. Isolation of a double-stranded RNA and a virus-like particle from Geotrichum candidum. J. Basic Microbiol. 1991, 31, 447–452. [Google Scholar] [CrossRef]
  255. Pospíšek, M.; Palková, Z.; Janderová, B.; Korb, J. Isolation and characterization of the dsRNA virus from the yeast Endomyces magnusii. FEMS Microbiol. Lett. 1994, 116, 231–236. [Google Scholar] [CrossRef] [PubMed]
  256. Groves, D.P.; Clare, J.J.; Oliver, S.G. Isolation and characterisation of a double-stranded RNA virus-like particle from the yeast Yarrowia lipolytica. Curr. Genet. 1983, 7, 185–190. [Google Scholar] [CrossRef] [PubMed]
  257. Abu-Mejdad, N.M.J.A.; Al-Badran, A.I.; Al-Saadoon, A.H.; Minati, M.H. A new report on gene expression of three killer toxin genes with antimicrobial activity of two killer toxins in Iraq. Bull. Natl. Res. Cent. 2020, 44, 162. [Google Scholar] [CrossRef]
  258. Lee, M.D.; Creagh, J.W.; Fredericks, L.R.; Crabtree, A.M.; Patel, J.S.; Rowley, P.A. The Characterization of a Novel Virus Discovered in the Yeast Pichia membranifaciens. Viruses 2022, 14, 594. [Google Scholar] [CrossRef]
  259. Forbes, E.M.; Nieduszynska, S.R.; Brunton, F.K.; Gibson, J.; Glover, L.A.; Stansfield, I. Control of gag-pol gene expression in the Candida albicans retrotransposon Tca2. BMC Mol. Biol. 2007, 8, 94. [Google Scholar] [CrossRef][Green Version]
  260. Schmitt, M.J.; Breinig, F. Yeast viral killer toxins: Lethality and self-protection. Nat. Rev. Microbiol. 2006, 4, 212–221. [Google Scholar] [CrossRef]
  261. Wickner, R.B.; Fujimura, T.; Esteban, R. Viruses and Prions of Saccharomyces cerevisiae. In Advances in Virus Research; Academic Press: New York, NY, USA, 2013; Volume 86, pp. 1–36. [Google Scholar]
  262. Rodríguez-Cousiño, N.; Gómez, P.; Esteban, R. Variation and distribution of LA helper totiviruses in Saccharomyces sensu stricto yeasts producing different killer toxins. Toxins 2017, 9, 313. [Google Scholar] [CrossRef][Green Version]
  263. Müller, F.; Brühl, K.-H.; Freidel, K.; Kowallik, K.V.; Ciriacy, M. Processing of TY1 proteins and formation of Ty1 virus-like particles in Saccharomyces cerevisiae. Mol. Gener. Genet. 1987, 207, 421–429. [Google Scholar] [CrossRef]
  264. Schmitt, M.J.; Neuhausen, F. Killer toxin-secreting double-stranded RNA mycoviruses in the yeasts Hanseniaspora uvarum and Zygosaccharomyces bailii. J. Virol. 1994, 68, 1765–1772. [Google Scholar] [CrossRef][Green Version]
  265. Depierreux, D.; Vong, M.; Nibert, M.L. Nucleotide sequence of Zygosaccharomyces bailii virus Z: Evidence for+ 1 programmed ribosomal frameshifting and for assignment to family Amalgaviridae. Virus Res. 2016, 217, 115–124. [Google Scholar] [CrossRef][Green Version]
  266. Schmitt, M.J.; Poravou, O.; Trenz, K.; Rehfeldt, K. Unique double-stranded RNAs responsible for the anti-Candida activity of the yeast Hanseniaspora uvarum. J. Virol. 1997, 71, 8852–8855. [Google Scholar] [CrossRef][Green Version]
  267. Pfeiffer, I.; Farkas, Z.; Golubev, W.I. dsRNA viruses in Nadsonia fulvescens. J. Gen. Appl. Microbiol. 2004, 50, 97–100. [Google Scholar] [CrossRef][Green Version]
  268. Crucitti, D.; Chiapello, M.; Oliva, D.; Forgia, M.; Turina, M.; Carimi, F.; La Bella, F.; Pacifico, D. Identification and molecular characterization of novel mycoviruses in Saccharomyces and non-Saccharomyces yeasts of oenological interest. Viruses 2021, 14, 52. [Google Scholar] [CrossRef]
  269. Tavantzis, S.M.; Romaine, C.P.; Smith, S.H. Purification and partial characterization of a bacilliform virus from Agaricus bisporus: A single-stranded RNA mycovirus. Virology 1980, 105, 94–102. [Google Scholar] [CrossRef]
  270. Kim, Y.-J.; Park, S.-H.; Yie, S.-W.; Kim, K.-H. RT-PCR detection of dsRNA Mycoviruses infecting Pleurotus ostreatus and Agaricus blazei Murrill. Plant Pathol. J. 2005, 21, 343–348. [Google Scholar] [CrossRef]
  271. Deakin, G.; Dobbs, E.; Bennett, J.M.; Jones, I.M.; Grogan, H.M.; Burton, K.S. Multiple viral infections in Agaricus bisporus-Characterisation of 18 unique RNA viruses and 8 ORFans identified by deep sequencing. Sci. Rep. 2017, 7, 2469. [Google Scholar] [CrossRef][Green Version]
  272. Heinze, C. A novel mycovirus from Clitocybe odora. Arch. Virol. 2012, 157, 1831–1834. [Google Scholar] [CrossRef]
  273. Shamoun, S.; Varga, A.; Valverde, R.; Ramsfield, T.; Sumampong, G.; Elliott, M.; Masri, S.; James, D. Identification and molecular characterization of a new double-stranded RNA virus infecting Chondrostereum purpureum. Can. J. Plant Pathol. 2008, 30, 604–613. [Google Scholar] [CrossRef]
  274. Sahin, E.; Akata, I. Full-length genome characterization of a novel alphapartitivirus detected in the ectomycorrhizal fungus Hygrophorus penarioides. Virus Genes 2021, 57, 94–99. [Google Scholar] [CrossRef]
  275. Sahin, E.; Akata, I.; Keskin, E. Molecular characterization of a new endornavirus inhabiting the ectomycorrhizal fungus Hygrophorus penarioides. Braz. J. Microbiol. 2021, 52, 1167–1172. [Google Scholar] [CrossRef] [PubMed]
  276. Akata, I.; Edis, G.; Keskin, E.; Sahin, E. Diverse partitiviruses hosted by the ectomycorrhizal agaric Hebeloma mesophaeum and the natural transmission of a partitivirus between phylogenetically distant, sympatric fungi. Virology 2023, 581, 63–70. [Google Scholar] [CrossRef]
  277. Sahin, E.; Keskin, E.; Akata, I. Molecular characterization of the complete genome of a novel partitivirus hosted by the saprobic mushroom Leucocybe candicans. Arch. Microbiol. 2021, 203, 5825–5830. [Google Scholar] [CrossRef] [PubMed]
  278. Reyes, B.M.D.; Fonseca, P.L.C.; Heming, N.M.; de Amorim Conceição, L.B.; de Souza Nascimento, K.T.; Gramacho, K.P.; Arevalo-Gardini, E.; Pirovani, C.P.; Aguiar, E.R.G.R. Characterization of the microbiota dynamics associated with Moniliophthora roreri, causal agent of cocoa frosty pod rot disease, reveals new viral species. Front. Microbiol. 2022, 13, 1053562. [Google Scholar] [CrossRef] [PubMed]
  279. Linnakoski, R.; Sutela, S.; Coetzee, M.P.; Duong, T.A.; Pavlov, I.N.; Litovka, Y.A.; Hantula, J.; Wingfield, B.D.; Vainio, E.J. Armillaria root rot fungi host single-stranded RNA viruses. Sci. Rep. 2021, 11, 7336. [Google Scholar] [CrossRef]
  280. Tonka, T.; Walterová, L.; Hejna, O.; Čurn, V. Molecular characterization of a ssRNA mycovirus isolated from the forest pathogenic fungus Armillaria ostoyae. Acta virologica 2022, 66, 290–294. [Google Scholar] [CrossRef]
  281. Zhao, Y.-j.; Shirouzu, T.; Chiba, Y.; Hosaka, K.; Moriyama, H.; Urayama, S.-i.; Hagiwara, D. Identification of novel RNA mycoviruses from wild mushroom isolates in Japan. Virus Res. 2023, 325, 199045. [Google Scholar] [CrossRef]
  282. Magae, Y.; Sunagawa, M. Characterization of a mycovirus associated with the brown discoloration of edible mushroom, Flammulina velutipes. Virol. J. 2010, 7, 342. [Google Scholar] [CrossRef][Green Version]
  283. Yu, H.J.; Lim, D.; Lee, H.-S. Characterization of a novel single-stranded RNA mycovirus in Pleurotus ostreatus. Virology 2003, 314, 9–15. [Google Scholar] [CrossRef][Green Version]
  284. Ro, H.-S.; Kang, E.-J.; Yu, J.-S.; Lee, T.-S.; Lee, C.-W.; Lee, H.-S. Isolation and characterization of a novel mycovirus, PeSV, in Pleurotus eryngii and the development of a diagnostic system for it. Biotechnol. Lett. 2007, 29, 129–135. [Google Scholar] [CrossRef]
  285. Li, X.; Li, S.; Yin, W.; Sossah, F.L.; Song, B.; Li, Y.; Zhang, B. Complete genome sequence of a novel mycovirus from Pleurotus citrinopileatus. Arch. Virol. 2023, 168, 66. [Google Scholar] [CrossRef]
  286. Chen, K.; Liang, P.; Yu, M.; Chang, S. A new double-stranded RNA virus from Volvariella volvacea. Mycologia 1988, 80, 849–853. [Google Scholar] [CrossRef]
  287. Barroso, G.; Labarère, J. Evidence for viral and naked double-stranded RNAs in the basidiomycete Agrocybe aegerita. Curr. Genet. 1990, 18, 231–237. [Google Scholar] [CrossRef]
  288. Li, X.; Sui, K.; Xie, J.; Hai, D.; Yin, W.; Sossah, F.L.; Jiang, D.; Song, B.; Li, Y. Molecular characterization of a novel fusarivirus infecting the edible fungus Auricularia heimuer. Arch. Virol. 2020, 165, 2689–2693. [Google Scholar] [CrossRef]
  289. Li, X.; Liu, Q.; Li, S.; Sossah, F.L.; Han, X.; Zhu, G.; Li, Y.; Li, C.; Fu, Y. A novel fungal negative-stranded RNA virus related to mymonaviruses in Auricularia heimuer. Arch. Virol. 2022, 167, 2223–2227. [Google Scholar] [CrossRef]
  290. Li, X.; Xie, J.; Hai, D.; Sui, K.; Yin, W.; Sossah, F.L.; Jiang, D.; Song, B.; Li, Y. Molecular characteristics of a novel ssRNA virus isolated from Auricularia heimuer in China. Arch. Virol. 2020, 165, 1495–1499. [Google Scholar] [CrossRef]
  291. Peter, M.; Courty, P.E.; Kohler, A.; Delaruelle, C.; Martin, D.; Tagu, D.; Frey-Klett, P.; Duplessis, S.; Chalot, M.; Podila, G. Analysis of expressed sequence tags from the ectomycorrhizal basidiomycetes Laccaria bicolor and Pisolithus microcarpus. New Phytol. 2003, 159, 117–129. [Google Scholar] [CrossRef]
  292. Cao, C.; Li, H.; Jones, M.G.; Wylie, S.J. Challenges to elucidating how endornaviruses influence fungal hosts: Creating mycovirus-free isogenic fungal lines and testing them. J. Virol. Methods 2019, 274, 113745. [Google Scholar] [CrossRef]
  293. Cao, C.T.; Derbyshire, M.C.; Regmi, R.; Li, H.; Jones, M.G.; Wylie, S.J. Small RNA Analyses of a Ceratobasidium Isolate Infected with Three Endornaviruses. Viruses 2022, 14, 2276. [Google Scholar] [CrossRef]
  294. Zhong, J.; Chen, C.-Y.; Gao, B.-D. Genome sequence of a novel mycovirus of Rhizoctonia solani, a plant pathogenic fungus. Virus Genes 2015, 51, 167–170. [Google Scholar] [CrossRef]
  295. Li, Y.; Sun, Y.; Yu, L.; Chen, W.; Liu, H.; Yin, L.; Guang, Y.; Yang, G.; Mo, X. Complete genome sequence of a novel mitovirus from binucleate Rhizoctonia AG-K strain FAS2909W. Arch. Virol. 2022, 167, 271–276. [Google Scholar] [CrossRef] [PubMed]
  296. Zheng, L.; Shu, C.; Zhang, M.; Yang, M.; Zhou, E. Molecular characterization of a novel endornavirus conferring hypovirulence in rice sheath blight fungus Rhizoctonia solani AG-1 IA strain GD-2. Viruses 2019, 11, 178. [Google Scholar] [CrossRef] [PubMed][Green Version]
  297. Abdoulaye, A.H.; Foda, M.F.; Kotta-Loizou, I. Viruses infecting the plant pathogenic fungus Rhizoctonia solani. Viruses 2019, 11, 1113. [Google Scholar] [CrossRef] [PubMed][Green Version]
  298. Marzano, S.-Y.L.; Domier, L.L. Novel mycoviruses discovered from metatranscriptomics survey of soybean phyllosphere phytobiomes. Virus Res. 2016, 213, 332–342. [Google Scholar] [CrossRef][Green Version]
  299. Komatsu, A.; Kondo, H.; Sato, M.; Kurahashi, A.; Nishibori, K.; Suzuki, N.; Fujimori, F. Isolation and characterization of a novel mycovirus infecting an edible mushroom, Grifola frondosa. Mycoscience 2019, 60, 211–220. [Google Scholar] [CrossRef]
  300. Kozlakidis, Z.; Hacker, C.V.; Bradley, D.; Jamal, A.; Phoon, X.; Webber, J.; Brasier, C.M.; Buck, K.W.; Coutts, R.H. Molecular characterisation of two novel double-stranded RNA elements from Phlebiopsis gigantea. Virus Genes 2009, 39, 132–136. [Google Scholar] [CrossRef]
  301. Drenkhan, T.; Sutela, S.; Veeväli, V.; Vainio, E.J. Phlebiopsis gigantea strains from Estonia show potential as native biocontrol agents against Heterobasidion root rot and contain diverse dsRNA and ssRNA viruses. Biol. Control. 2022, 167, 104837. [Google Scholar] [CrossRef]
  302. Akata, I.; Keskin, E.; Sahin, E. Molecular characterization of a new mitovirus hosted by the ectomycorrhizal fungus Albatrellopsis flettii. Arch. Virol. 2021, 166, 3449–3454. [Google Scholar] [CrossRef]
  303. Vainio, E.J.; Sutela, S. Mixed infection by a partitivirus and a negative-sense RNA virus related to mymonaviruses in the polypore fungus Bondarzewia berkeleyi. Virus Res. 2020, 286, 198079. [Google Scholar] [CrossRef]
  304. Sutela, S.; Piri, T.; Vainio, E.J. Discovery and community dynamics of novel ssRNA mycoviruses in the conifer pathogen Heterobasidion parviporum. Front. Microbiol. 2021, 12, 770787. [Google Scholar] [CrossRef]
  305. Vainio, E.J.; Korhonen, K.; Tuomivirta, T.T.; Hantula, J. A novel putative partitivirus of the saprotrophic fungus Heterobasidion ecrustosum infects pathogenic species of the Heterobasidion annosum complex. Fungal Biol. 2010, 114, 955–965. [Google Scholar] [CrossRef]
  306. Vainio, E.J.; Hantula, J. Taxonomy, biogeography and importance of Heterobasidion viruses. Virus Res. 2016, 219, 2–10. [Google Scholar] [CrossRef]
  307. Petrzik, K.; Sarkisova, T.; Starý, J.; Koloniuk, I.; Hrabáková, L.; Kubešová, O. Molecular characterization of a new monopartite dsRNA mycovirus from mycorrhizal Thelephora terrestris (Ehrh.) and its detection in soil oribatid mites (Acari: Oribatida). Virology 2016, 489, 12–19. [Google Scholar] [CrossRef][Green Version]
  308. Teng, L.; Li, X.; Cai, X.; Yang, S.; Liu, H.; Zhang, T. The complete genome sequence of a novel mycovirus in the plant-pathogenic fungus Exobasidium gracile. Arch. Virol. 2022, 167, 1343–1347. [Google Scholar] [CrossRef]
  309. Zhang, T.; Cai, X.; Teng, L.; Li, X.; Zhong, N.; Liu, H. Molecular characterization of three novel mycoviruses in the plant pathogenic fungus Exobasidium. Virus Res. 2022, 307, 198608. [Google Scholar] [CrossRef]
  310. Beck, R.J.; Smith, O.P.; Tooley, P.W.; Peterson, G.L.; Bonde, M.R. Characterization of double-stranded RNA from Tilletia indica. Mycologia 1994, 86, 656–659. [Google Scholar] [CrossRef]
  311. Applen Clancey, S.; Ruchti, F.; LeibundGut-Landmann, S.; Heitman, J.; Ianiri, G. A novel mycovirus evokes transcriptional rewiring in the fungus Malassezia and stimulates beta interferon production in macrophages. Mbio 2020, 11, e01534-20. [Google Scholar] [CrossRef]
  312. Zhang, T.; Cai, X.; Zeng, Z.; Du, H.; Liu, H. A novel totivirus isolated from the phytopathogenic fungus Rhodosporidiobolus odoratus strain GZ2017. Arch. Virol. 2020, 165, 1911–1914. [Google Scholar] [CrossRef]
  313. Liu, J.-J.; Xiang, Y.; Sniezko, R.A.; Schoettle, A.W.; Williams, H.; Zamany, A. Characterization of Cronartium ribicola dsRNAs reveals novel members of the family Totiviridae and viral association with fungal virulence. Virol. J. 2019, 16, 118. [Google Scholar] [CrossRef]
  314. Liu, J.-J.; Chan, D.; Xiang, Y.; Williams, H.; Li, X.-R.; Sniezko, R.A.; Sturrock, R.N. Characterization of five novel mitoviruses in the white pine blister rust fungus Cronartium ribicola. PLoS ONE 2016, 11, e0154267. [Google Scholar] [CrossRef][Green Version]
  315. Lawrence, G.; Boelen, M.; Pryor, A. Transmission of double-stranded RNAs in flax rust, Melampsora lini. Canad. J. Bot. 1988, 66, 61–66. [Google Scholar] [CrossRef]
  316. Zheng, L.; Zhao, J.; Liang, X.; Zhuang, H.; Qi, T.; Kang, Z. Complete genome sequence of a novel mitovirus from the wheat stripe rust fungus Puccinia striiformis. Arch. Virol. 2019, 164, 897–901. [Google Scholar] [CrossRef] [PubMed]
  317. Pryor, A.; Boelen, M. A double-stranded RNA mycovirus from the maize rust Puccinia sorghi. Canad. J. Bot. 1987, 65, 2380–2383. [Google Scholar] [CrossRef]
  318. Zheng, L.; Lu, X.; Liang, X.; Jiang, S.; Zhao, J.; Zhan, G.; Liu, P.; Wu, J.; Kang, Z. Molecular characterization of novel totivirus-like double-stranded RNAs from Puccinia striiformis f. sp. tritici, the causal agent of wheat stripe rust. Front. Microbiol. 2017, 8, 1960. [Google Scholar] [CrossRef][Green Version]
  319. Dickinson, M.J.; Pryor, A. Isometric virus-like particles encapsidate the double-stranded RNA found in Puccinia striiformis, Puccinia recondita, and Puccinia sorghi. Canad. J. Bot. 1989, 67, 3420–3425. [Google Scholar] [CrossRef]
  320. Castillo, A.; Cifuentes, V. Presence of double-stranded RNA and virus-like particles in Phaffia rhodozyma. Curr. Genet. 1994, 26, 364–368. [Google Scholar] [CrossRef]
  321. Baeza Cancino, M.; Bravo, N.; Sanhueza, M.; Flores, O.; Villarreal, P.; Cifuentes Guzmán, V. Molecular characterization of totiviruses in Xanthophyllomyces dendrorhous. Virol. J. 2012, 9, 140. [Google Scholar] [CrossRef][Green Version]
  322. Golubev, W.I.; Pfeiffer, I.; Churkina, L.G.; Golubeva, E.W. Double-stranded RNA viruses in a mycocinogenic strain of Cystofilobasidium infirmominiatum. FEMS Yeast Res. 2003, 3, 63–68. [Google Scholar] [CrossRef]
  323. Gácser, A.; Hamari, Z.; Pfeiffer, I.; Varga, J.; Kevei, F.; Kucsera, J. Genetic diversity in the red yeast Cryptococcus hungaricus and its phylogenetic relationship to some related basidiomycetous yeasts. FEMS Yeast Res. 2001, 1, 213–220. [Google Scholar]
  324. Golubev, W.I.; Pfeiffer, I.; Golubeva, E. Mycocin production in Trichosporon pullulans populations colonizing tree exudates in the spring. FEMS Microbiol. Ecol. 2002, 40, 151–157. [Google Scholar] [CrossRef]
  325. Khandjian, E.; Turian, G.; Eisen, H. Characterization of the RNA mycovirus infecting Allomyces arbuscula. J. Gen. Virol. 1977, 35, 415–424. [Google Scholar] [CrossRef]
  326. Khandjian, E.; Turian, G. In vitro RNA synthesis by double-stranded RNA mycovirus from Allomyces arbuscula. FEMS Microbiol. Lett. 1977, 2, 121–124. [Google Scholar] [CrossRef]
  327. Papp, T.; Nyilasi, I.; Fekete, C.; Ferenczy, L.; Vágvölgyi, C. Presence of double-stranded RNA and virus-like particles in Rhizopus isolates. Can. J. Microbiol. 2001, 47, 443–447. [Google Scholar] [CrossRef]
  328. Turina, M.; Ghignone, S.; Astolfi, N.; Silvestri, A.; Bonfante, P.; Lanfranco, L. The virome of the arbuscular mycorrhizal fungus Gigaspora margarita reveals the first report of DNA fragments corresponding to replicating non-retroviral RNA viruses in fungi. Environ. Microbiol. 2018, 20, 2012–2025. [Google Scholar] [CrossRef]
  329. Ikeda, Y.; Shimura, H.; Kitahara, R.; Masuta, C.; Ezawa, T. A novel virus-like double-stranded RNA in an obligate biotroph arbuscular mycorrhizal fungus: A hidden player in mycorrhizal symbiosis. Mol. Plant Microbe Interact. 2012, 25, 1005–1012. [Google Scholar] [CrossRef][Green Version]
  330. Kitahara, R.; Ikeda, Y.; Shimura, H.; Masuta, C.; Ezawa, T. A unique mitovirus from Glomeromycota, the phylum of arbuscular mycorrhizal fungi. Arch. Virol. 2014, 159, 2157–2160. [Google Scholar] [CrossRef][Green Version]
  331. Kartali, T.; Nyilasi, I.; Kocsubé, S.; Patai, R.; Polgár, T.F.; Zsindely, N.; Nagy, G.; Bodai, L.; Lipinszki, Z.; Vágvölgyi, C. Characterization of four novel dsRNA viruses isolated from Mucor hiemalis strains. Viruses 2021, 13, 2319. [Google Scholar] [CrossRef]
  332. Barata, R.R.; Vianez-Júnior, J.L.; Nunes, M.R. Transcriptomic analysis of Mucor irregularis containing a negative single-stranded RNA mycovirus. Microbiol. Resour. Announc. 2019, 8, e00503-19. [Google Scholar] [CrossRef][Green Version]
  333. Kartali, T.; Nyilasi, I.; Szabó, B.; Kocsubé, S.; Patai, R.; Polgár, T.F.; Nagy, G.; Vágvölgyi, C.; Papp, T. Detection and molecular characterization of novel dsRNA viruses related to the Totiviridae family in Umbelopsis ramanniana. Front. Cell. Infect. Microbiol. 2019, 9, 249. [Google Scholar] [CrossRef][Green Version]
  334. Qu, J.; Shi, N.; Yang, G.; Huang, B. Molecular characterization of a novel totivirus infecting the basal fungus Conidiobolus heterosporus. Arch. Virol. 2021, 166, 1801–1804. [Google Scholar] [CrossRef]
  335. Shi, N.; Xie, T.; Yang, G.; Wang, P.; Huang, B. Molecular characterization of two novel totiviruses coinfecting the basal fungus Conidiobolus adiaeretus. Arch. Virol. 2023, 168, 7. [Google Scholar] [CrossRef] [PubMed]
  336. Yang, G.; Hu, F.; Shi, N.; Wang, P.; Huang, B. A novel non-segmented double-stranded RNA virus isolated from the basal fungus Conidiobolus sp. Arch. Virol. 2020, 165, 1919–1923. [Google Scholar] [CrossRef] [PubMed]
  337. Krupovic, M.; Dolja, V.V.; Koonin, E.V. Plant viruses of the Amalgaviridae family evolved via recombination between viruses with double-stranded and negative-strand RNA genomes. Biol. Direct 2015, 10, 12. [Google Scholar] [CrossRef] [PubMed][Green Version]
  338. Yang, Y.; Gaspard, G.; McMullen, N.; Duncan, R. Polycistronic Genome Segment Evolution and Gain and Loss of FAST Protein Function during Fusogenic Orthoreovirus Speciation. Viruses 2020, 12, 702. [Google Scholar] [CrossRef] [PubMed]
  339. Vainio, E.J.; Chiba, S.; Ghabrial, S.A.; Maiss, E.; Roossinck, M.; Sabanadzovic, S.; Suzuki, N.; Xie, J.; Nibert, M. ICTV Virus Taxonomy Profile: Partitiviridae. J. Gen. Virol. 2018, 99, 17–18. [Google Scholar] [CrossRef]
  340. Chiba, S.; Castón, J.R.; Ghabrial, S.A.; Suzuki, N. ICTV Virus Taxonomy Profile: Quadriviridae. J. Gen. Virol. 2018, 99, 1480–1481. [Google Scholar] [CrossRef]
  341. Sato, Y.; Miyazaki, N.; Kanematsu, S.; Xie, J.; Ghabrial, S.A.; Hillman, B.I.; Suzuki, N. ICTV Virus Taxonomy Profile: Megabirnaviridae. J. Gen. Virol. 2019, 100, 1269–1270. [Google Scholar] [CrossRef]
  342. Kotta-Loizou, I.; Castón, J.R.; Coutts, R.H.A.; Hillman, B.I.; Jiang, D.; Kim, D.-H.; Moriyama, H.; Suzuki, N. ICTV Virus Taxonomy Profile: Chrysoviridae. J. Gen. Virol. 2020, 101, 143–144. [Google Scholar] [CrossRef]
  343. Kotta-Loizou, I.; Coutts, R.H.; Consortium, I.R. ICTV Virus Taxonomy Profile: Polymycoviridae 2022. J. Gen. Virol. 2022, 103, 1747. [Google Scholar] [CrossRef]
  344. Supyani, S.; Hillman, B.I.; Suzuki, N. Baculovirus expression of the 11 mycoreovirus-1 genome segments and identification of the guanylyltransferase-encoding segment. J. Gen. Virol. 2007, 88, 342–350. [Google Scholar] [CrossRef]
  345. Lin, Y.H.; Hisano, S.; Yaegashi, H.; Kanematsu, S.; Suzuki, N. A second quadrivirus strain from the phytopathogenic filamentous fungus Rosellinia necatrix. Arch. Virol. 2013, 158, 1093–1098. [Google Scholar] [CrossRef]
  346. Li, L.; Liu, J.; Xu, A.; Wang, T.; Chen, J.; Zhu, X. Molecular characterization of a trisegmented chrysovirus isolated from the radish Raphanus sativus. Virus Res. 2013, 176, 169–178. [Google Scholar] [CrossRef]
  347. Valverde, R.A.; Khalifa, M.E.; Okada, R.; Fukuhara, T.; Sabanadzovic, S. ICTV Virus Taxonomy Profile: Endornaviridae. J. Gen. Virol. 2019, 100, 1204–1205. [Google Scholar] [CrossRef]
  348. Gilbert, K.B.; Holcomb, E.E.; Allscheid, R.L.; Carrington, J.C. Discovery of new mycoviral genomes within publicly available fungal transcriptomic datasets. PLoS ONE 2019, 14, e0219207. [Google Scholar]
  349. Li, P.; Zhang, H.; Chen, X.; Qiu, D.; Guo, L. Molecular characterization of a novel hypovirus from the plant pathogenic fungus Fusarium graminearum. Virology 2015, 481, 151–160. [Google Scholar] [CrossRef][Green Version]
  350. Park, D.; Goh, C.; Lee, J.; Sebastiani, F.; Hahn, Y. Identification of Pistacia-associated flexivirus 1, a putative mycovirus of the family Gammaflexiviridae, in the mastic tree (Pistacia lentiscus) transcriptome. Acta Virol. 2020, 64, 28–35. [Google Scholar] [CrossRef]
  351. Chen, X.; He, H.; Yang, X.; Zeng, H.; Qiu, D.; Guo, L. The complete genome sequence of a novel Fusarium graminearum RNA virus in a new proposed family within the order Tymovirales. Arch. Virol. 2016, 161, 2899–2903. [Google Scholar] [CrossRef]
  352. Jiāng, D.; Ayllón, M.A.; Marzano, S.-Y.L. ICTV virus taxonomy profile: Mymonaviridae. J. Gen. Virol. 2019, 100, 1343–1344. [Google Scholar] [CrossRef]
  353. Wang, Y.; Zhao, H.; Cao, J.; Yin, X.; Guo, Y.; Guo, L.; Wu, H.; Zhang, M. Characterization of a Novel Mycovirus from the Phytopathogenic Fungus Botryosphaeria dothidea. Viruses 2022, 14, 331. [Google Scholar] [CrossRef]
  354. Llorens, C.; Soriano, B.; Krupovic, M. ICTV Virus Taxonomy Profile: Metaviridae. J. Gen. Virol. 2020, 101, 1131–1132. [Google Scholar] [CrossRef]
  355. Llorens, C.; Soriano, B.; Krupovic, M. ICTV Virus Taxonomy Profile: Pseudoviridae. J. Gen. Virol. 2021, 102, 1563. [Google Scholar] [CrossRef] [PubMed]
  356. Dodonova, S.O.; Prinz, S.; Bilanchone, V.; Sandmeyer, S.; Briggs, J.A. Structure of the Ty3/Gypsy retrotransposon capsid and the evolution of retroviruses. Proc. Natl. Acad. Sci. USA 2019, 116, 10048–10057. [Google Scholar] [CrossRef] [PubMed][Green Version]
  357. 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] [PubMed]
  358. Tran, T.; Li, H.; Nguyen, D.; Jones, M.; Wylie, S. Co-Infection with Three Mycoviruses Stimulates Growth of a Monilinia fructicola Isolate on Nutrient Medium, but Does Not Induce Hypervirulence in a Natural Host. Viruses 2019, 11, 89. [Google Scholar] [CrossRef][Green Version]
  359. Nuss, D.L. Mycoviruses, RNA Silencing, and Viral RNA Recombination. In Advances in Virus Research; Maramorosch, K., Shatkin, A.J., Murphy, F.A., Eds.; Academic Press: New York, NY, USA, 2011; Volume 80, pp. 25–48. [Google Scholar]
  360. Zhang, D.-X.; Spiering, M.J.; Nuss, D.L. Characterizing the Roles of Cryphonectria parasitica RNA-Dependent RNA Polymerase-Like Genes in Antiviral Defense, Viral Recombination and Transposon Transcript Accumulation. PLoS ONE 2014, 9, e108653. [Google Scholar] [CrossRef][Green Version]
  361. Segers, G.C.; Zhang, X.; Deng, F.; Sun, Q.; Nuss, D.L. Evidence that RNA silencing functions as an antiviral defense mechanism in fungi. Proc. Natl. Acad. Sci. USA 2007, 104, 12902–12906. [Google Scholar] [CrossRef][Green Version]
  362. Sun, Q.; Choi, G.H.; Nuss, D.L. A single Argonaute gene is required for induction of RNA silencing antiviral defense and promotes viral RNA recombination. Proc. Natl. Acad. Sci. USA 2009, 106, 17927–17932. [Google Scholar] [CrossRef][Green Version]
  363. Chang, S.-S.; Zhang, Z.; Liu, Y. RNA Interference Pathways in Fungi: Mechanisms and Functions. Annu. Rev. Microbiol. 2012, 66, 305–323. [Google Scholar] [CrossRef][Green Version]
  364. Zhao, J.H.; Guo, H.S. RNA silencing: From discovery and elucidation to application and perspectives. J. Integr. Plant Biol. 2022, 64, 476–498. [Google Scholar] [CrossRef]
  365. Segers, G.C.; van Wezel, R.; Zhang, X.; Hong, Y.; Nuss, D.L. Hypovirus papain-like protease p29 suppresses RNA silencing in the natural fungal host and in a heterologous plant system. Eukaryot. Cell. 2006, 5, 896–904. [Google Scholar] [CrossRef][Green Version]
  366. Aulia, A.; Hyodo, K.; Hisano, S.; Kondo, H.; Hillman, B.I.; Suzuki, N. Identification of an RNA Silencing Suppressor Encoded by a Symptomless Fungal Hypovirus, Cryphonectria Hypovirus 4. Biology 2021, 10, 100. [Google Scholar] [CrossRef]
  367. Yaegashi, H.; Yoshikawa, N.; Ito, T.; Kanematsu, S. A mycoreovirus suppresses RNA silencing in the white root rot fungus, Rosellinia necatrix. Virology 2013, 444, 409–416. [Google Scholar] [CrossRef][Green Version]
  368. Gaitanis, G.; Magiatis, P.; Hantschke, M.; Bassukas, I.D.; Velegraki, A. The Malassezia Genus in Skin and Systemic Diseases. Clin. Microbiol. Rev. 2012, 25, 106–141. [Google Scholar] [CrossRef][Green Version]
  369. Limon, J.J.; Tang, J.; Li, D.; Wolf, A.J.; Michelsen, K.S.; Funari, V.; Gargus, M.; Nguyen, C.; Sharma, P.; Maymi, V.I. Malassezia is associated with Crohn’s disease and exacerbates colitis in mouse models. Cell Host Microbe 2019, 25, 377–388.e6. [Google Scholar] [CrossRef][Green Version]
  370. Aykut, B.; Pushalkar, S.; Chen, R.; Li, Q.; Abengozar, R.; Kim, J.I.; Shadaloey, S.A.; Wu, D.; Preiss, P.; Verma, N. The fungal mycobiome promotes pancreatic oncogenesis via activation of MBL. Nature 2019, 574, 264–267. [Google Scholar] [CrossRef]
  371. Kotta-Loizou, I.; Sipkova, J.; Coutts, R.H. Identification and sequence determination of a novel double-stranded RNA mycovirus from the entomopathogenic fungus Beauveria bassiana. Arch. Virol. 2015, 160, 873–875. [Google Scholar] [CrossRef]
  372. Chen, B.; Gao, S.; Choi, G.H.; Nuss, D.L. Extensive alteration of fungal gene transcript accumulation and elevation of G-protein-regulated cAMP levels by a virulence-attenuating hypovirus. Proc. Natl. Acad. Sci. USA 1996, 93, 7996–8000. [Google Scholar] [CrossRef] [PubMed][Green Version]
  373. Choi, G.H.; Chen, B.; Nuss, D.L. Virus-mediated or transgenic suppression of a G-protein alpha subunit and attenuation of fungal virulence. Proc. Natl. Acad. Sci. USA 1995, 92, 305–309. [Google Scholar] [CrossRef][Green Version]
  374. Gao, S.; Nuss, D.L. Distinct roles for two G protein α subunits in fungal virulence, morphology, and reproduction revealed by targeted gene disruption. Proc. Natl. Acad. Sci. USA 1996, 93, 14122–14127. [Google Scholar] [CrossRef][Green Version]
  375. Craven, M.; Pawlyk, D.; Choi, G.; Nuss, D. Papain-like protease p29 as a symptom determinant encoded by a hypovirulence-associated virus of the chestnut blight fungus. J. Virol. 1993, 67, 6513–6521. [Google Scholar] [CrossRef][Green Version]
  376. Shi, L.; Wang, J.; Quan, R.; Yang, F.; Shang, J.; Chen, B. CpATG8, a homolog of yeast autophagy protein ATG8, is required for pathogenesis and hypovirus accumulation in the chest blight fungus. Front. Cell. Infect. Microbiol. 2019, 9, 222. [Google Scholar] [CrossRef] [PubMed]
  377. Ko, Y.-H.; So, K.-K.; Chun, J.; Kim, D.-H. Distinct roles of two DNA methyltransferases from Cryphonectria parasitica in fungal virulence, responses to hypovirus infection, and viral clearance. MBio 2021, 12, e02890-20. [Google Scholar] [CrossRef] [PubMed]
  378. Chu, Y.-M.; Jeon, J.-J.; Yea, S.-J.; Kim, Y.-H.; Yun, S.-H.; Lee, Y.-W.; Kim, K.-H. Double-stranded RNA mycovirus from Fusarium graminearum. Appl. Environ. Microbiol. 2002, 68, 2529–2534. [Google Scholar] [CrossRef][Green Version]
  379. Lemus-Minor, C.G.; Cañizares, M.C.; García-Pedrajas, M.D.; Pérez-Artés, E. Horizontal and vertical transmission of the hypovirulence-associated mycovirus Fusarium oxysporum f. sp. dianthi virus 1. Eur. J. Plant Pathol. 2019, 153, 645–650. [Google Scholar] [CrossRef]
  380. Sasaki, A.; Kanematsu, S.; Onoue, M.; Oyama, Y.; Yoshida, K. Infection of Rosellinia necatrix with purified viral particles of a member of Partitiviridae (RnPV1-W8). Arch. Virol. 2006, 151, 697–707. [Google Scholar] [CrossRef]
  381. Ran, H.; Liu, L.; Li, B.; Cheng, J.; Fu, Y.; Jiang, D.; Xie, J. Co-infection of a hypovirulent isolate of Sclerotinia sclerotiorum with a new botybirnavirus and a strain of a mitovirus. Virol. J. 2016, 13, 92. [Google Scholar] [CrossRef][Green Version]
  382. Urayama, S.-i.; Sakoda, H.; Takai, R.; Katoh, Y.; Le, T.M.; Fukuhara, T.; Arie, T.; Teraoka, T.; Moriyama, H. A dsRNA mycovirus, Magnaporthe oryzae chrysovirus 1-B, suppresses vegetative growth and development of the rice blast fungus. Virology 2014, 448, 265–273. [Google Scholar] [CrossRef][Green Version]
  383. Wu, M.; Zhang, L.; Li, G.; Jiang, D.; Hou, M.; Huang, H.-C. Hypovirulence and double-stranded RNA in Botrytis cinerea. Phytopathology 2007, 97, 1590–1599. [Google Scholar] [CrossRef][Green Version]
  384. Wu, M.; Zhang, L.; Li, G.; Jiang, D.; Ghabrial, S.A. Genome characterization of a debilitation-associated mitovirus infecting the phytopathogenic fungus Botrytis cinerea. Virology 2010, 406, 117–126. [Google Scholar] [CrossRef][Green Version]
  385. Zhong, J.; Li, P.; Gao, B.D.; Zhong, S.Y.; Li, X.G.; Hu, Z.; Zhu, J.Z. Novel and diverse mycoviruses co-infecting a single strain of the phytopathogenic fungus Alternaria dianthicola. Front. Cell. Infect. Microbiol. 2022, 12, 1427. [Google Scholar] [CrossRef]
  386. Tebbi, C.K.; Badiga, A.; Sahakian, E.; Powers, J.J.; Achille, A.N.; Patel, S.; Migone, F. Exposure to a mycovirus containing Aspergillus Flavus reproduces acute lymphoblastic leukemia cell surface and genetic markers in cells from patients in remission and not controls. Cancer Treat. Res. Commun. 2021, 26, 100279. [Google Scholar] [CrossRef] [PubMed]
  387. Rowley, P.A. The frenemies within: Viruses, retrotransposons and plasmids that naturally infect Saccharomyces yeasts. Yeast 2017, 34, 279–292. [Google Scholar] [CrossRef] [PubMed][Green Version]
  388. 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]
  389. Schmitt, M.J.; Klavehn, P.; Wang, J.; Schnig, I.; Tipper, D.J. Cell cycle studies on the mode of action of yeast K28 killer toxin. Microbiology 1996, 142, 2655–2662. [Google Scholar] [CrossRef][Green Version]
  390. Martinac, B.; Zhu, H.; Kubalski, A.; Zhou, X.; Culbertson, M.; Bussey, H.; Kung, C. Yeast K1 killer toxin forms ion channels in sensitive yeast spheroplasts and in artificial liposomes. Proc. Natl. Acad. Sci. USA 1990, 87, 6228–6232. [Google Scholar] [CrossRef][Green Version]
  391. Orentaite, I.; Poranen, M.M.; Oksanen, H.M.; Daugelavicius, R.; Bamford, D.H. K2 killer toxin-induced physiological changes in the yeast Saccharomyces cerevisiae. FEMS Yeast Res. 2016, 16, fow003. [Google Scholar] [CrossRef][Green Version]
  392. Liu, S.-Q.; Tsao, M. Inhibition of spoilage yeasts in cheese by killer yeast Williopsis saturnus var. saturnus. Int. J. Food Microbiol. 2009, 131, 280–282. [Google Scholar] [CrossRef]
  393. 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]
  394. Hutzler, M.; Riedl, R.; Koob, J.; Jacob, F. Fermentation and spoilage yeasts and their relevance for the beverage industry-a review. Brew Sci. 2012, 65, 33–52. [Google Scholar]
  395. Mehlomakulu, N.; Setati, M.; Divol, B. Non-Saccharomyces killer toxins: Possible biocontrol agents against Brettanomyces in wine? S. Afr. J. Enol. Vitic. 2015, 36, 94–104. [Google Scholar] [CrossRef][Green Version]
  396. Santos, A.; Navascués, E.; Bravo, E.; Marquina, D. Ustilago maydis killer toxin as a new tool for the biocontrol of the wine spoilage yeast Brettanomyces bruxellensis. Int. J. Food Microbiol. 2011, 145, 147–154. [Google Scholar] [CrossRef] [PubMed]
  397. Weiler, F.; Schmitt, M.J. Zygocin, a secreted antifungal toxin of the yeast Zygosaccharomyces bailii, and its effect on sensitive fungal cells. FEMS Yeast Res. 2003, 3, 69–76. [Google Scholar] [CrossRef] [PubMed]
  398. Bissegger, M.; Rigling, D.; Heiniger, U. Population Structure and Disease Development of Cryphonectria parasitica in European Chestnut Forests in the Presence of Natural Hypovirulence. Phytopathology 1997, 87, 50–59. [Google Scholar] [CrossRef][Green Version]
  399. Heiniger, U.; Rigling, D. Biological control of chestnut blight in Europe. Annu. Rev. Phytopathol. 1994, 32, 581–599. [Google Scholar] [CrossRef]
  400. Rigling, D.; Prospero, S. Cryphonectria parasitica, the causal agent of chestnut blight: Invasion history, population biology and disease control. Mol. Plant Pathol. 2018, 19, 7–20. [Google Scholar] [CrossRef][Green Version]
  401. Milgroom, M.G.; Cortesi, P. Biological control of chestnut blight with hypovirulence: A critical analysis. Annu. Rev. Phytopathol. 2004, 42, 311–338. [Google Scholar] [CrossRef][Green Version]
  402. Heiniger, U.; Rigling, D. Application of the cryphonectria hypovirus (chv-1) to control the chestnut blight, experience from switzerland. Acta Hortic. 2009, 815, 233–246. [Google Scholar] [CrossRef]
  403. Prospero, S.; Rigling, D. Using molecular markers to assess the establishment and spread of a mycovirus applied as a biological control agent against chestnut blight. BioControl 2016, 61, 313–323. [Google Scholar] [CrossRef]
  404. Yu, X.; Li, B.; Fu, Y.; Xie, J.; Cheng, J.; Ghabrial, S.A.; Li, G.; Yi, X.; Jiang, D. Extracellular transmission of a DNA mycovirus and its use as a natural fungicide. Proc. Natl. Acad. Sci. USA 2013, 110, 1452–1457. [Google Scholar] [CrossRef][Green Version]
  405. Refos, J.M.; Vonk, A.G.; Eadie, K.; Lo-Ten-Foe, J.R.; Verbrugh, H.A.; Van Diepeningen, A.D.; Van De Sande, W.W.J. Double-Stranded RNA Mycovirus Infection of Aspergillus fumigatus Is Not Dependent on the Genetic Make-Up of the Host. PLoS ONE 2013, 8, e77381. [Google Scholar] [CrossRef][Green Version]
  406. Glass, N.L.; Jacobson, D.J.; Shiu, P.K.T. The genetics of hyphal fusion and vegetative incompatibility in filamentous ascomycete fungi. Annu. Rev. Genet. 2000, 34, 165–186. [Google Scholar] [CrossRef] [PubMed][Green Version]
  407. García-Pedrajas, M.; Cañizares, M.; Sarmiento-Villamil, J.L.; Jacquat, A.G.; Dambolena, J.S. Mycoviruses in biological control: From basic research to field implementation. Phytopathology 2019, 109, 1828–1839. [Google Scholar] [CrossRef] [PubMed]
  408. Anagnostakis, S.L. Biological control of chestnut blight. Science 1982, 215, 466–471. [Google Scholar] [CrossRef] [PubMed]
  409. Attanayake, R.N.; Carter, P.A.; Jiang, D.; del Río-Mendoza, L.; Chen, W. Sclerotinia sclerotiorum populations infecting canola from China and the United States are genetically and phenotypically distinct. Phytopathology 2013, 103, 750–761. [Google Scholar] [CrossRef] [PubMed][Green Version]
  410. Kohn, L.M.; Carbone, I.; Anderson, J.B. Mycelial interactions in Sclerotinia sclerotiorum. Exp. Mycol. 1990, 14, 255–267. [Google Scholar] [CrossRef]
  411. Zhang, D.-X.; Nuss, D.L. Engineering super mycovirus donor strains of chestnut blight fungus by systematic disruption of multilocus vic genes. Proc. Natl. Acad. Sci. USA 2016, 113, 2062–2067. [Google Scholar] [CrossRef][Green Version]
  412. Hutchison, E.; Brown, S.; Tian, C.; Glass, N.L. Transcriptional profiling and functional analysis of heterokaryon incompatibility in Neurospora crassa reveals that reactive oxygen species, but not metacaspases, are associated with programmed cell death. Microbiology 2009, 155, 3957–3970. [Google Scholar] [CrossRef][Green Version]
  413. Ikeda, K.; Inoue, K.; Kida, C.; Uwamori, T.; Sasaki, A.; Kanematsu, S.; Park, P. Potentiation of mycovirus transmission by zinc compounds via attenuation of heterogenic incompatibility in Rosellinia necatrix. Appl. Environ. Microbiol. 2013, 79, 3684–3691. [Google Scholar] [CrossRef][Green Version]
  414. Brusini, J.; Robin, C. Mycovirus transmission revisited by in situ pairings of vegetatively incompatible isolates of Cryphonectria parasitica. J. Virol. Methods 2013, 187, 435–442. [Google Scholar] [CrossRef]
  415. Yaegashi, H.; Nakamura, H.; Sawahata, T.; Sasaki, A.; Iwanami, Y.; Ito, T.; Kanematsu, S. Appearance of mycovirus-like double-stranded RNAs in the white root rot fungus, Rosellinia necatrix, in an apple orchard. FEMS Microbiol. Ecol. 2013, 83, 49–62. [Google Scholar] [CrossRef][Green Version]
  416. Bryner, S.F.; Rigling, D. Virulence not only costs but also benefits the transmission of a fungal virus. Evolution 2012, 66, 2540–2550. [Google Scholar] [CrossRef]
  417. Biella, S.; Smith, M.L.; Aist, J.R.; Cortesi, P.; Milgroom, M.G. Programmed cell death correlates with virus transmission in a filamentous fungus. Proc. R. Soc. Lond. B Biol. Sci. 2002, 269, 2269–2276. [Google Scholar] [CrossRef]
  418. Shang, J.; Wu, X.; Lan, X.; Fan, Y.; Dong, H.; Deng, Y.; Nuss, D.L.; Chen, B. Large-scale expressed sequence tag analysis for the chestnut blight fungus Cryphonectria parasitica. Fungal Genet. Biol. 2008, 45, 319–327. [Google Scholar] [CrossRef]
  419. Wu, S.; Cheng, J.; Fu, Y.; Chen, T.; Jiang, D.; Ghabrial, S.A.; Xie, J. Virus-mediated suppression of host non-self recognition facilitates horizontal transmission of heterologous viruses. PLoS Pathog. 2017, 13, e1006234. [Google Scholar] [CrossRef][Green Version]
  420. Sande, W.W.J.; Lo-Ten-Foe, J.R.; Belkum, A.; Netea, M.G.; Kullberg, B.J.; Vonk, A.G. Mycoviruses: Future therapeutic agents of invasive fungal infections in humans? Eur. J. Clin. Microbiol. Infect. Dis. 2010, 29, 755–763. [Google Scholar] [CrossRef][Green Version]
  421. Marzano, S.-Y.L.; Hobbs, H.A.; Nelson, B.D.; Hartman, G.L.; Eastburn, D.M.; McCoppin, N.K.; Domier, L.L. Transfection of Sclerotinia sclerotiorum with in vitro transcripts of a naturally occurring interspecific recombinant of Sclerotinia sclerotiorum hypovirus 2 significantly reduces virulence of the fungus. J. Virol. 2015, 89, 5060–5071. [Google Scholar] [CrossRef][Green Version]
  422. Chen, B.; Choi, G.H.; Nuss, D.L. Attenuation of fungal virulence by synthetic infectious hypovirus transcripts. Science 1994, 264, 1762–1764. [Google Scholar] [CrossRef]
  423. Choi, G.H.; Nuss, D.L. Hypovirulence of chestnut blight fungus conferred by an infectious viral cDNA. Science 1992, 257, 800–803. [Google Scholar] [CrossRef]
  424. Sasaki, A.; Onoue, M.; Kanematsu, S.; Suzaki, K.; Miyanishi, M.; Suzuki, N.; Nuss, D.L.; Yoshida, K. Extending Chestnut Blight Hypovirus Host Range Within Diaporthales by Biolistic Delivery of Viral cDNA. Mol. Plant Microbe Interact. 2002, 15, 780–789. [Google Scholar] [CrossRef][Green Version]
  425. Hillman, B.I.; Supyani, S.; Kondo, H.; Suzuki, N. A reovirus of the fungus Cryphonectria parasitica that is infectious as particles and related to the Coltivirus genus of animal pathogens. J. Virol. 2004, 78, 892–898. [Google Scholar] [CrossRef][Green Version]
  426. Clausen, M.; Kräuter, R.; Schachermayr, G.; Potrykus, I.; Sautter, C. Antifungal activity of a virally encoded gene in transgenic wheat. Nat. Biotechnol. 2000, 18, 446–449. [Google Scholar] [CrossRef] [PubMed]
  427. Polashock, J.; Bedker, P.; Hillman, B. Movement of a small mitochondrial double-stranded RNA element of Cryphonectria parasitica: Ascospore inheritance and implications for mitochondrial recombination. Mol. Gener. Genet. 1997, 256, 566–571. [Google Scholar] [CrossRef] [PubMed]
  428. Van Diepeningen, A.D.; Debets, A.J.; Hoekstra, R.F. Dynamics of dsRNA mycoviruses in black Aspergillus populations. Fungal Genet. Biol. 2006, 43, 446–452. [Google Scholar] [CrossRef] [PubMed]
  429. Yang, D.; Wu, M.; Zhang, J.; Chen, W.; Li, G.; Yang, L. Sclerotinia minor Endornavirus 1, a Novel Pathogenicity Debilitation-Associated Mycovirus with a Wide Spectrum of Horizontal Transmissibility. Viruses 2018, 10, 589. [Google Scholar] [CrossRef][Green Version]
  430. Hillman, B.I.; Suzuki, N. Viruses in the chestnut blight fungus. In Advances in Virus Research; Academic Press: New York, NY, USA, 2004; Volume 63, pp. 423–473. [Google Scholar]
  431. Lee, K.-M.; Cho, W.K.; Yu, J.; Son, M.; Choi, H.; Min, K.; Lee, Y.-W.; Kim, K.-H. A Comparison of Transcriptional Patterns and Mycological Phenotypes following Infection of Fusarium graminearum by Four Mycoviruses. PLoS ONE 2014, 9, e100989. [Google Scholar] [CrossRef]
  432. Ihrmark, K.; Stenström, E.; Stenlid, J. Double-stranded RNA transmission through basidiospores of Heterobasidion annosum. Mycol. Res. 2004, 108, 149–153. [Google Scholar] [CrossRef]
  433. Varga, J.; Rinyu, E.; Kevei, É.; Tóth, B.; Kozakiewicz, Z. Double-stranded RNA mycoviruses in species of Aspergillus sections Circumdati and Fumigati. Can. J. Microbiol. 1998, 44, 569–574. [Google Scholar] [CrossRef]
  434. Morris, T.; Dodds, J. Isolation and analysis of double-stranded RNA from virus-infected plant and fungal tissue. Phytopathology 1979, 69, 854–858. [Google Scholar] [CrossRef][Green Version]
  435. Okada, R.; Kiyota, E.; Moriyama, H.; Fukuhara, T.; Natsuaki, T. A simple and rapid method to purify viral dsRNA from plant and fungal tissue. J. Gen. Plant Pathol. 2015, 81, 103–107. [Google Scholar] [CrossRef]
  436. Dawe, A.L.; Nuss, D.L. Hypovirus Molecular Biology: From Koch’s Postulates to Host Self-Recognition Genes that Restrict Virus Transmission. In Advances in Virus Research; Ghabrial, S.A., Ed.; Academic Press: New York, NY, USA, 2013; Volume 86, pp. 109–147. [Google Scholar]
  437. Liu, R.; Cheng, J.; Fu, Y.; Jiang, D.; Xie, J. Molecular Characterization of a Novel Positive-Sense, Single-Stranded RNA Mycovirus Infecting the Plant Pathogenic Fungus Sclerotinia sclerotiorum. Viruses 2015, 7, 2470–2484. [Google Scholar] [CrossRef][Green Version]
  438. Xie, J.; Xiao, X.; Fu, Y.; Liu, H.; Cheng, J.; Ghabrial, S.A.; Li, G.; Jiang, D. A novel mycovirus closely related to hypoviruses that infects the plant pathogenic fungus Sclerotinia sclerotiorum. Virology 2011, 418, 49–56. [Google Scholar] [CrossRef][Green Version]
Figure 1. Mycoviral taxa currently recognized by the ICTV. Compiled using information from the official ICTV website (https://talk.ictvonline.org/, (accessed on 27 March 2023)).
Figure 1. Mycoviral taxa currently recognized by the ICTV. Compiled using information from the official ICTV website (https://talk.ictvonline.org/, (accessed on 27 March 2023)).
Viruses 15 01202 g001
Table 1. Mycoviruses associated with Ascomycota.
Table 1. Mycoviruses associated with Ascomycota.
Ascomycota
ClassOrderFamilyGeneraMycoviruses
ArthoniomycetesArthonialesChrysotrichaceaeChrysothrixdsRNA [42]
DothideomycetesBotryosphaerialesAplosporellaceaeAplosporelladsRNA [43]
BotryosphaeriaceaeBotryosphaeriadsRNA and +ssRNA [44,45,46,47,48]
DiplodiadsRNA and +ssRNA [49,50,51]
Eutiarosporella+ssRNA [52]
MacrophominadsRNA, +ssRNA, −ssRNA and ssDNA [53,54,55,56]
NeofusicoccumdsRNA and +ssRNA [57,58,59]
PhaeobotryondsRNA [60]
PhyllostictaceaePhyllostictadsRNA [61]
CapnodialesCladosporiaceaeCladosporium+ssRNA, −ssRNA and RT ssRNA [62,63]
MycosphaerellaceaeCercospora+ssRNA and −ssRNA [64]
DothistromadsRNA [65]
Mycosphaerella+ssRNA [52,66]
PseudocercosporadsRNA [66]
ZymoseptoriadsRNA and +ssRNA [61,64,67]
DothidealesSaccotheciaceaeAureobasidium+ssRNA [52]
MycosphaerellalesTeratosphaeriaceaeAcidomyces+ssRNA [61]
HortaeadsRNA [61]
PleosporalesConiothyriaceaeConiothyriumdsRNA and −ssRNA [63,68]
CorynesporascaceaeCorynesporadsRNA [67,69]
CucurbitariaceaeCucurbitariadsRNA [69,70]
DelitschiaceaeDelitschiadsRNA [61]
DidymellaceaeDidymelladsRNA [71]
Epicoccum+ssRNA [63]
Leptosphaerulina+ssRNA [63]
PhomadsRNA and +ssRNA [71,72,73,74,75]
StagonosporopsisdsRNA [47,72,73,74,75]
LeptosphaeriaceaeLeptosphaeriadsRNA and + ssRNA [76,77,78]
LindgomycetaceaeClohesyomycesdsRNA [47,61]
MassarinaceaeHelminthosporiumdsRNA [79]
PericoniaceaePericoniadsRNA and +ssRNA [61,80]
PleosporaceaeAlternariadsRNA, +ssRNA, and −ssRNA [63,80,81,82,83,84,85,86,87,88]
BipolarisdsRNA and +ssRNA [89,90,91,92]
CochliobolusdsRNA [93]
CurvulariadsRNA [94]
DrechsleradsRNA [61,95]
ExserohilumdsRNA [95,96,97]
PleosporadsRNA [96,97,98]
SetosphaeriadsRNA and +ssRNA [97,99,100]
Stemphylium+ssRNA [101]
EurotiomycetesChaetothyrialesHerpotrichiellaceaePhialophoradsRNA [81,82,83,84,85,86,87,88,102,103]
EurotialesAspergillaceaeAspergillusdsRNA and +ssRNA [93,104,105,106,107]
PenicilliumdsRNA, +ssRNA and −ssRNA [15,63,89,90,91,92,108,109]
TrichocomaceaeTalaromycesdsRNA [94,110,111,112]
OnygenalesAjellomycetaceaeBlastomycesdsRNA [98,113]
AscosphaeraceaeAscosphaera+ssRNA [97,99,100,114,115]
Onygenales incertae sedisMyriodontiumdsRNA [116]
LecanoromycetesLecanoralesStereocaulaceaeLeprariadsRNA [42,101]
SordariomycetesDiaporthalesCryphonectriaceaeCryphonectriadsRNA, +ssRNA and −ssRNA [117,118,119,120,121]
EndothiadsRNA [122]
DiaporthaceaeDiaporthedsRNA and +ssRNA [56,123,124]
MelanconiellaceaeMelanconiella+ssRNA [125]
ValsaceaeCytosporadsRNA [126]
PhomopsisdsRNA and +ssRNA [127,128,129,130]
Valsa+ssRNA [131]
GlomerellalesGlomerellaceaeColletotrichumdsRNA and +ssRNA [132,133,134,135,136,137,138]
PlectosphaerellaceaeSodiomycesdsRNA and +ssRNA [139]
VerticilliumdsRNA and +ssRNA [140,141,142,143]
HypocrealesBionectriaceaeClonostachysdsRNA [144]
ClavicipitaceaeAtkinsonelladsRNA [145]
EpichloëdsRNA [146,147]
MetarhiziumdsRNA [148,149,150,151]
UstilaginoideadsRNA [152,153,154,155,156]
CordycipitaceaeBeauveriadsRNA [61,157,158,159,160]
CordycepsdsRNA [161,162]
IsariadsRNA [163]
HypocreaceaeHypomycesdsRNA [164]
TrichodermadsRNA and +ssRNA [61,165,166,167,168]
FusariumdsRNA, +ssRNA, −ssRNA, and ssDNA [169,170,171,172,173,174,175]
IlyonectriadsRNA [176,177]
NectriadsRNA [178]
RugonectriadsRNA and +ssRNA [179]
ThelonectriadsRNA [180]
Hypocreales incertae sedisAcremonium+ssRNA [63]
OphiocordycipitaceaeOphiocordyceps+ssRNA [61]
TolypocladiumdsRNA [181,182]
MagnaporthalesMagnaporthaceaeBuergenerula+ssRNA [183]
GaeumannomycesdsRNA and +ssRNA [61,184]
MagnaporthedsRNA and +ssRNA [185,186,187,188]
MicroascalesCeratocystidaceaeBerkeleyomycesdsRNA and +ssRNA [189,190]
CeratocystisdsRNA and +ssRNA [191]
EndoconidiophoradsRNA [192]
PyriculariaceaePyriculariaceaedsRNA [193,194]
OphiostomatalesOphiostomataceaeGrosmanniadsRNA and +ssRNA [61]
OphiostomadsRNA and +ssRNA [195,196,197]
SordarialesSordariaceaeNeurosporadsRNA and +ssRNA [198]
TogninialesTogniniaceaePhaeoacremoniumdsRNA and +ssRNA [63]
XylarialesApiosporaceaeNigrosporadsRNA and +ssRNA [199,200,201,202]
DiatrypaceaeMonosporascusdsRNA [203]
HypoxylaceaeAnnulohypoxylon+ssRNA [204]
MicrodochiaceaeMicrodochium+ssRNA [205]
SporocadaceaePestalotiopsisdsRNA, +ssRNA and −ssRNA [206]
PseudopestalotiopsisdsRNA [207]
XylariaceaeEntoleucadsRNA and +ssRNA [208,209,210]
RoselliniadsRNA and +ssRNA [18,208,211,212,213,214,215,216,217]
LeotiomycetesLeotiomycetes incertae sedisMyxotrichaceaeOidiodendron+ssRNA [70,218]
PseudeurotiaceaePseudogymnoascusdsRNA [219]
ErysiphalesErysiphaceaeErysiphedsRNA and +ssRNA [68,220,221]
PodosphaeradsRNA and +ssRNA [77]
HelotialesGodroniaceaeGremmenielladsRNA and +ssRNA [51,61,222,223]
HelotiaceaeHymenoscyphusdsRNA, +ssRNA and −ssRNA [121,224,225]
MollisiaceaeLoramyces+ssRNA [61]
RutstroemiaceaeRutstroemia+ssRNA [61]
SclerotiniaceaeBotrytisdsRNA and +ssRNA [226,227,228,229,230,231]
MoniliniadsRNA, +ssRNA and ssDNA [232,233]
SclerotiniadsRNA, +ssRNA, −ssRNA and dsRNA [16,175,234,235,236,237,238]
SclerotiumdsRNA and +ssRNA [239,240]
ThelebolalesThelebolaceaeThelebolusdsRNA [61]
PezizomycetesPezizalesCaloscyphaceaeCaloscyphadsRNA [241]
DiscinaceaeGyromitradsRNA [61,242]
MorchellaceaeMorchelladsRNA and +ssRNA [61,243]
PezizaceaePeziza+ssRNA [204]
SarcosphaeradsRNA [244]
TerfeziadsRNA [245]
PyronemataceaeGeoporadsRNA and +ssRNA [121,246]
PicoadsRNA and +ssRNA [247]
TuberaceaeTuberdsRNA, +ssRNA and RT-ssRNA [121,248,249,250]
SaccharomycetesSaccharomycetalesDebaryomycetaceaeScheffersomycesdsRNA [251]
WickerhamiadsRNA [252]
DipodascaceaeGeotrichumdsRNA [253,254]
MagnusiomycesdsRNA [255]
YarrowiadsRNA [256]
PhaffomycetaceaeWickerhamomycesdsRNA [257]
PichiaceaePichiadsRNA [258]
SaccharomycetaceaeCandidaRT-ssRNA [259]
SaccharomycesdsRNA, +ssRNA and RT-ssRNA [260,261,262,263]
TorulasporadsRNA [257]
ZygosaccharomycesdsRNA [264,265]
SaccharomycodaceaeHanseniasporadsRNA [264,266]
Saccharomycetales incertae sedisAmbrosiozymadsRNA [258]
NadsoniadsRNA [267]
StarmerelladsRNA [268]
MagnaporthedsRNA and +ssRNA [185,186,187,188]
Table 2. Mycoviruses associated with Basidiomycota.
Table 2. Mycoviruses associated with Basidiomycota.
Basidiomycota
ClassOrderFamilyGeneraMycoviruses
AgaricomycetesAgaricalesAgaricaceaeAgaricusdsRNA and +ssRNA [269,270,271]
Leucocoprinus+ssRNA [52]
ClitocybaceaeClitocybe+ssRNA [272]
CyphellaceaeChondrostereumdsRNA [204,273]
HydnangiaceaeLaccaria+ssRNA [204]
HygrophoraceaeHygrophorusdsRNA [121,274,275]
HymenogastraceaeHebelomadsRNA [276]
LyophyllaceaeLeucocybedsRNA [277]
MarasmiaceaeMoniliophthora+ssRNA [278]
NidulariaceaeCyathus+ssRNA [52]
OmphalotaceaeCollybiopsis+ssRNA [204]
LentinuladsRNA and +ssRNA [204]
PhysalacriaceaeArmillaria+ssRNA and −ssRNA [121,279,280]
FlammulinadsRNA [281,282]
PleurotaceaePleurotusdsRNA and +ssRNA [270,283,284,285]
PluteaceaeVolvarielladsRNA [204,286]
PorotheleaceaeMegacollybiadsRNA [204]
PsathyrellaceaeCoprinopsidsRNA and +ssRNA [204,281]
SchizophyllaceaeSchizophyllum+ssRNA [204]
StrophariaceaeAgrocybedsRNA [281,287]
SquamanitaceaePhaeolepiota+ssRNA [204]
AuricularialesAuriculariaceaeAuriculariadsRNA, +ssRNA and −ssRNA [288,289,290]
BoletalesBoletaceaeBoletus+ssRNA [204]
PisolithaceaePisolithusdsRNA [291]
CantharellalesCantharellaceaeCraterellus+ssRNA [204]
CeratobasidiaceaeCeratobasidium+ssRNA [292,293]
RhizoctoniadsRNA, +ssRNA and −ssRNA [294,295,296,297]
Thanatephorus+ssRNA [298]
TulasnellaceaeTulasnella−ssRNA [218]
GloeophyllalesGloeophyllaceaeNeolentinus+ssRNA [204]
HymenochaetalesHymenochaetaceaeFomitiporia+ssRNA [63]
PolyporalesGrifolaceaeGrifoladsRNA [299]
PhanerochaetaceaePhlebiopsisdsRNA and +ssRNA [300,301]
RussulalesAlbatrellaceaeAlbatrellopsis+ssRNA [302]
BondarzewiaceaeBondarzewiadsRNA and −ssRNA [303]
HeterobasidiondsRNA and +ssRNA [304,305,306]
RussulaceaeLactariusdsRNA [121]
ThelephoralesThelephoraceaeThelephoradsRNA [204,307]
ExobasidiomycetesExobasidialesExobasidiaceaeExobasidiumdsRNA [304,308,309]
TilletialesTilletiaceaeTilletiadsRNA [310]
MalasseziomycetesMalassezialesMalasseziaceaeMalasseziadsRNA [311]
MicrobotryomycetesSporidiobolalesSporidiobolaceaeRhodosporidiobolusdsRNA [312]
PuccinialesColeosporiaceaeCronartiumdsRNA and +ssRNA [121,313,314]
MelampsoraceaeMelampsoradsRNA [315]
PucciniaceaePucciniadsRNA and +ssRNA [316,317,318,319]
UromycesdsRNA [52]
PhakopsoraceaePhakopsoradsRNA [52]
TremellomycetesCystofilobasidialesMrakiaceaePhaffiadsRNA [320,321]
CystofilobasidiaceaeCystofilobasidiumdsRNA [322]
TremellalesCryptococcaceaeCryptococcusdsRNA [323]
TrichosporonalesTrichosporonaceaeTrichosporondsRNA [324]
WallemiomycetesWallemialesWallemiaceaeWallemiadsRNA [98]
Table 3. Mycoviruses associated with the early diverging lineages of fungi.
Table 3. Mycoviruses associated with the early diverging lineages of fungi.
Blastocladiomycota
ClassOrderFamilyGeneraMycoviruses
BlastocladiomycetesBlastocladialesBlastocladiaceaeAllomycesdsRNA [2,325,326]
Chytridiomycota
ClassOrderFamilyGeneraMycoviruses
ChytridiomycetesChytridialesChytridiaceaeZopfochytriumdsRNA [2]
CladochytrialesCladochytriaceaeCladochytriumdsRNA [2]
RhizophydialesRhizophydiales incertae sedisOperculomycesdsRNA and +ssRNA [2]
RhizopodaceaeRhizopusdsRNA and +ssRNA [2,327]
Mucoromycota
ClassOrderFamilyGeneraMycoviruses
GlomeromycetesArchaeosporalesGeosiphonaceaeGeosiphon+ssRNA [2]
DiversisporalesGigasporaceaeGigasporadsRNA, +ssRNA and ssDNA [2,52,328]
GlomeralesGlomeraceaeGlomusdsRNA [329]
RacocetradsRNA and +ssRNA [52]
Rhizophagus+ssRNA [2,32,330]
ParaglomeralesParaglomeraceaeParaglomus+ ssRNA [52]
MortierellomycetesMortierellalesMortierellaceaeDissophoradsRNA [2]
LobosporangiumdsRNA [2]
MortierelladsRNA, +ssRNA and −ssRNA [2]
MucoromycetesMucoralesChoanephoraceaeBlakesleadsRNA [2]
ChoanephoradsRNA and +ssRNA [2]
CunninghamellaceaeAbsidiadsRNA [2]
MucoraceaeMucordsRNA, −ssRNA and ssDNA [52,53,331,332]
PhycomycetaceaePhycomyces[2]
SyncephalastraceaeSyncephalastrumdsRNA [107]
UmbelopsidomycetesUmbelopsidalesUmbelopsidaceaeUmbelopsisdsRNA [2,41,333]
Neocallimastigomycota
ClassOrderFamilyGeneraMycoviruses
NeocallimastigomycetesNeocallimastigalesNeocallimastigaceaeAnaeromyces+ssRNA [2]
NeocallimastixdsRNA [2]
PecoramycesssDNA [52]
Zoopagomycota
ClassOrderFamilyGeneraMycoviruses
EntomophthoromycetesEntomophthoralesAncylistaceaeConidiobolusdsRNA and +ssRNA [2,334,335,336]
EntomophthoraceaeEntomophagadsRNA [2]
Entomophthora+ssRNA [2,52]
ZoophthoradsRNA and +ssRNA [2]
KickxellomycetesKickxellalesKickxellaceaeKickxelladsRNA and +ssRNA [2]
ZoopagomycetesZoopagalesPiptocephalidaceaeSyncephalis+ssRNA [2]
Table 5. Genomic features and characteristics of + ssRNA mycoviruses.
Table 5. Genomic features and characteristics of + ssRNA mycoviruses.
Families and GeneraGenome Size and SegmentationGenome OrganizationExemplar Species
Family
Alphaflexiviridae
Mycovirus associated genera
Botrexvirus
Sclerodarnavirus
Non-segmented genomes around 5.5 to 9 kb in size.Botrexvirus:
Genome consists of up to 5 ORFs.
ORF1: Encodes for an RdRp, Hel, and Mtf.
ORF3: Encodes for a CP.
All remaining ORFs encode for putative proteins with unknown function.
Sclerodarnavirus:
Genome consists of a single ORF encoding for an RdRp, Mtf, and Hel.
Botrexvirus:
Botrytis virus X
Sclerodarnavirus:
Sclerotinia sclerotiorum debilitation-associated RNA virus
Family
Barnaviridae
Mycovirus associated genera
Barnavirus
Non-segmented with genomes around 4.0 kb in size.Genome consists of 4 ORFs.
ORF1: Encodes for a hypothetical protein with an unknown function.
ORF2: Encodes for a putative serine protease.
ORF3: Encodes for an RdRp.
ORF4: Encodes for a CP.
Mushroom bacilliform virus
Family
Botourmiaviridae
Mycovirus associated genera
Botoulivirus
Betabotoulivirus
Magoulivirus
Scleroulivirus
Betascleroulivirus
Deltascleroulivirus
Penoulivirus
Rhizoulivirus
Betarhizoulivirus
Non-segmented genomes around 3 to 5.3 kbp in sizeGenome consists of a single ORF encoding for an RdRpBotoulivirus:
Botrytis cinerea ourmia-like virus 4
Betabotoulivirus:
Entoleuca ourmia-like virus 1
Magoulivirus:
Magnaporthe oryzae ourmia-like virus 1
Scleroulivirus:
Sclerotinia sclerotiorum ourmia-like virus 1
Betascleroulivirus:
Botrytis cinerea ourmia-like virus 10
Deltascleroulivirus:
Botrytis cinerea ourmia-like virus 5
Penoulivirus:
Aspergillus neoniger ourmia-like virus 1
Rhizoulivirus:
Rhizoctonia solani ourmia-like virus 1Rs
Betarhizoulivirus:
Rhizoctonia solani ourmia-like virus 5
Family
Deltaflexiviridae
Mycovirus associated genera
Deltaflexivirus
Non-segmented with genomes around 6 to 8 kbp in sizeGenome consists of 1–5 ORFs
ORF 1: Encodes for polyprotein with Mt, Gtf, Hel and RdRp domains
ORF 2–5: Encode for hypothetical proteins with unknown functions
Sclerotinia sclerotiorum deltaflexivirus 1
Family
Endornaviridae
Mycovirus associated genera
Alphaendornavirus
Betaendornavirus
Non segmented genomes around 9.5 to 17.6 kb in size.Genome consists of a single ORF encoding for a large polyprotein.
The polyprotein always contains an RdRp domain, but may also include domains for Hel, Mt, Gtf, CPS, and phytoreo S7 domains.
Alphaendornavirus:
Oryza sativa alphaendornavirus
Betaendornavirus:
Sclerotinia sclerotiorum endornavirus
Family
Gammaflexiviridae
Mycovirus associated genera
Mycoflexivirus
Non-segmented genomes around 6.8 to 9.2 kb in size.Consists of 2–3 ORFs
ORF 1: Encodes for a replicase (REP) with an upstream Mtf domain, a Hel domain, and a downstream RdRp domain.
ORF 2: Encodes for either a CP or proteins which resemble movement proteins.
ORF 3: Encodes for either a CP or a hypothetical protein with an unknown function.
Botrytis virus F
Family
Hadakaviridae
Mycovirus associated genera
Hadakavirus
Multisegmented genomes around 14 to 15 kb in total size.RNA 1: Encodes for an RdRp.
RNA3: Encodes for a Mtf.
RNA 2, 4–7, 9 and 10: Encode for hypothetical proteins with an unknown function.
RNA8: May encode for a C2H2-type zinc finger protein
hadaka virus 1
Family
Hypoviridae
Mycovirus associated genera
Alphahypovirus
Betahypovirus
Epsilonhypovirus
Etahypovirus
Thetahypovirus
Non-segmented genomes around 9.1 to 12.7 kb in sizeConsist of 1–2 ORFs
Encodes for a polyprotein containing RdRp and sometimes protease, Hel, and Gtf domains.
Some hypoviruses also have short, internally deleted, defective interfering replicative forms of dsRNA molecules, while others have replicative forms of satellite like RNAs
Alphahypovirus
Cryphonectria hypovirus 2
Betahypovirus:
Cryphonectria hypovirus 4
Epsilonhypovirus:
Agaricus bisporus virus 2
Etahypovirus:
Sclerotium rolfsii hypovirus 8
Thetahypovirus:
Botrytis cinerea hypovirus 4
Family
Mitoviridae
Mycovirus associated genera
Unuamitovirus
Duamitovirus
Triamitovirus
Kvaramitovirus
Non-segmented genomes around 2.0 kbp to 4.5 kb in sizeConsists of 1 ORF encoding for an RdRpUnuamitovirus:
Botrytis cinerea mitovirus 2
Duamitovirus:
Alternaria alternata mitovirus 1
Triamitovirus:
Rhizoctonia solani mitovirus 30
Kvaramitovirus:
Ophiostoma mitovirus 7
Family
Narnaviridae
Mycovirus associated genera
Narnavirus
Non-segmented genomes around 2.3 to 3.6 kb in sizeConsists of 1 ORF encoding for an RdRpSaccharomyces 20S RNA narnavirus
Family
Yadokariviridae
Mycovirus associated genera
Alphayadokarivirus
Betayadokarivirus
Non-segmented genomes around 3.6 to 6.3 kb in sizeGenome consists of a 1–2 ORFs encoding for a polyprotein.
Alphayadokarivirus
Consist of 1 ORF which encodes for a polyprotein with an RdRp domain, 2A-like self-cleaving peptide and a hypothetical protein domain with an unknown function.
Betayadokarivirus
Consist of 1 or 2 ORFs encoding for a polyprotein, which may or may not include a 2A-like self-cleaving peptide. The polyprotein contains an RdRp domain (5′ proximal ORF), and a hypothetical protein domain (3′ proximal ORF) with an unknown function.
yado-kari virus 2
Compiled using data from the International Committee on Taxonomy of Viruses (ICTV): https://ictv.global/taxonomy/ (accessed on 27 March 2023). Hamid, et al. [234], Li, Zheng, Cheng, Chen, Fu, Jiang and Xie [235].), Ma, Zhang, Qi, Zhang, Ma, Jiang, Qin and Qi [172], Li, Sun, Yu, Chen, Liu, Yin, Guang, Yang and Mo [295].
Table 6. Genomic features and characteristics of—ssRNA mycoviruses.
Table 6. Genomic features and characteristics of—ssRNA mycoviruses.
Families and GeneraGenome Size and SegmentationGenome OrganizationExemplar Species
Family
Mymonaviridae
Mycovirus associated genera
Auricularimonavirus
Botrytimonavirus
Lentimonavirus
Penicillimonavirus
Sclerotimonavirus
Non-segmented genomes around 6 to 10 kb in sizeConsists of 1–7 ORFs
All mymonaviruses encode for an RdRp near the C-terminus, and some may encode for a NP near the 5′ terminus. Some also encode for hypothetical proteins with unknown functions.
Auricularimonavirus:
Auricularia heimuer negative-stranded RNA virus 1
Botrytimonavirus:
Botrytis cinerea negative-stranded RNA virus 5
Lentimonavirus:
Lentinula edodes negative-strand RNA virus 1
Penicillimonavirus:
Penicillium adametzioides negative-stranded RNA virus 1
Sclerotimonavirus:
Botrytis cinerea negative-stranded RNA virus 3
Compiled using data from the International Committee on Taxonomy of Viruses (ICTV): https://ictv.global/taxonomy/ (accessed on 20 March 2023).
Table 7. Genomic features and characteristics of RT ssRNA mycoviruses.
Table 7. Genomic features and characteristics of RT ssRNA mycoviruses.
Families and GeneraGenome Size and SegmentationGenome OrganizationExemplar Species
Family
Metaviridae
Mycovirus associated genera
Metavirus
Non-segmented genomes around 3 to 15 kb in sizeGenome consists of an intragenic region which may code for up to 2 genes (gag and pol), which is flanked by long terminal repeat sequences (LTR)
The gag gene usually encodes for the CP or NC protein while the pol gene encodes a polyprotein with PR, RT, INT and RH domains
Cladosporium fulvum T-1 virus
Family
Pseudoviridae
Mycovirus associated genera
Hemivirus
Pseudovirus
Non-segmented genomes that range in length from 4 to 9 kb in sizeGenome consists of an internal region with 1 (gag-pol) or 2 ORFs (gag and pol), which are flanked by long terminal repeat sequences (LTR)
The gag protein typically contains the CP and NC domains, while the pol protein contains PR, RT, INT, and RH domains
Hemivirus:
Candida albicans Tca2 virus
Pseudovirus:
Saccharomyces cerevisiae Ty1 virus
Compiled using data from the International Committee on Taxonomy of Viruses (ICTV): https://ictv.global/taxonomy/ (accessed on 20 March 2023).
Table 8. Genomic features and characteristics of ssDNA mycoviruses.
Table 8. Genomic features and characteristics of ssDNA mycoviruses.
Families and GeneraGenome Size and SegmentationGenome OrganizationExemplar Species
Family
Genomoviridae
Mycovirus associated genera
Gemycircularvirus
Gemytripvirus
Gemycircularvirus
Monopartite genomes around 1.8 to 2.4 kb in size
Gemytripvirus
Tripartite genome with individual segments around 1.3 kb
Gemycircularvirus
Genome encodes for a rolling circle REP protein as well as a CP in ambisense orientation
Gemytripvirus
DNA-A: Encodes for the REP protein
DNA-B: Encodes for the CP
DNA-C: Encodes for p26, a protein with an unknown function
Gemycircularvirus
Sclerotinia sclerotiorum hypovirulence-associated DNA virus 1
Gemytripvirus
Fusarium graminearum gemytripvirus 1
Compiled using data from the International Committee on Taxonomy of Viruses (ICTV): https://ictv.global/taxonomy/ (accessed on 20 March 2023), Li, Wang, Zhang, Qiu, Zhou and Guo [175] and Varsani and Krupovic [14].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hough, B.; Steenkamp, E.; Wingfield, B.; Read, D. Fungal Viruses Unveiled: A Comprehensive Review of Mycoviruses. Viruses 2023, 15, 1202. https://doi.org/10.3390/v15051202

AMA Style

Hough B, Steenkamp E, Wingfield B, Read D. Fungal Viruses Unveiled: A Comprehensive Review of Mycoviruses. Viruses. 2023; 15(5):1202. https://doi.org/10.3390/v15051202

Chicago/Turabian Style

Hough, Bianca, Emma Steenkamp, Brenda Wingfield, and David Read. 2023. "Fungal Viruses Unveiled: A Comprehensive Review of Mycoviruses" Viruses 15, no. 5: 1202. https://doi.org/10.3390/v15051202

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