Next Article in Journal
Dynamics and Patterning of 5-Hydroxytryptamine 2 Subtype Receptors in JC Polyomavirus Entry
Next Article in Special Issue
Canker Development and Biocontrol Potential of CHV-1 Infected English Isolates of Cryphonectria parasitica Is Dependent on the Virus Concentration and the Compatibility of the Fungal Inoculums
Previous Article in Journal
Current Epidemiology and Co-Infections of Avian Immunosuppressive and Neoplastic Diseases in Chicken Flocks in Central China
Previous Article in Special Issue
Hypovirulence of Colletotrichum gloesporioides Associated with dsRNA Mycovirus Isolated from a Mango Orchard in Thailand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bunyaviruses Affect Growth, Sporulation, and Elicitin Production in Phytophthora cactorum

1
Natural Resources Institute Finland (Luke), Latokartanonkaari 9, FI-00790 Helsinki, Finland
2
Phytophthora Research Centre, Department of Forest Protection and Wildlife Management, Faculty of Forestry and Wood Technology, Mendel University in Brno, Zemědělská 3, 613 00 Brno, Czech Republic
3
Natural Resources Institute Finland, Toivonlinnantie 518, FI-21500 Piikkiö, Finland
4
Phytophthora Research Centre, Department of Molecular Biology and Radiobiology, Faculty of AgriSciences, Mendel University in Brno, Zemědělská 1, 613 00 Brno, Czech Republic
5
Natural Resources Institute Finland, Humppilantie 18, FI-31600 Jokioinen, Finland
*
Author to whom correspondence should be addressed.
Viruses 2022, 14(12), 2596; https://doi.org/10.3390/v14122596
Submission received: 23 September 2022 / Revised: 2 November 2022 / Accepted: 20 November 2022 / Published: 22 November 2022
(This article belongs to the Special Issue Viruses and Their Effects on Fungal Host Fitness)

Abstract

:
Phytophthora cactorum is an important oomycetous plant pathogen with numerous host plant species, including garden strawberry (Fragaria × ananassa) and silver birch (Betula pendula). P. cactorum also hosts mycoviruses, but their phenotypic effects on the host oomycete have not been studied earlier. In the present study, we tested polyethylene glycol (PEG)-induced water stress for virus curing and created an isogenic virus-free isolate for testing viral effects in pair with the original isolate. Phytophthora cactorum bunya-like viruses 1 and 2 (PcBV1 & 2) significantly reduced hyphal growth of the P. cactorum host isolate, as well as sporangia production and size. Transcriptomic and proteomic analyses revealed an increase in the production of elicitins due to bunyavirus infection. However, the presence of bunyaviruses did not seem to alter the pathogenicity of P. cactorum. Virus transmission through anastomosis was unsuccessful in vitro.

1. Introduction

The oomycete genus Phytophthora includes notorious plant pathogens that severely threaten agricultural crops and forest trees worldwide. Phytophthora cactorum is an omnivorous pathogen that can infect a wide variety of plant species [1]. On cultivated strawberry, it is the causative agent of crown rot as well as leather rot of fruits. Both diseases cause economic losses in strawberry production globally [1,2]. The degree of resistance to the disease varies among strawberry cultivars [3]. Crown and leather rots are controlled mainly by propagating plants for planting from healthy stocks and by planting strawberries on ridges to improve drainage and aeration (hilling), but soil sanitation and fungicides are also used [4].
Phytophthora cactorum is a genetically variable species with two cosmopolitan lineages associated with strawberry and woody species (apple, oak, ornamental trees), respectively [5]. Finnish isolates from birch have been inferred to represent the most early-diverging lineage within the species and separate phylogenetically from the two cosmopolitan lineages [5]. Isolates from woody species, including Finnish isolates from birch, have caused little or no crown rot symptoms when inoculated in strawberry, but strawberry isolates have caused necrotic lesions on B. pendula [6,7,8]. Phytophthora cactorum can persist in the soil or symptomless plants for planting as sexual oospores or chlamydospores that germinate under wet conditions to produce mycelium and sporangia, which release motile zoospores to infect plant crowns or fruits [9].
Fungi and oomycetes host viruses that cause persistent, usually asymptomatic, infections in their host. Viruses reported from Phytophthora species include many alphaendornaviruses infecting P. cactorum, Phytophthora ramorum, a yet unnamed species designated as Phytophthora taxon douglas fir, and a Phytophthora sp. from asparagus [10,11,12]. Even single host species may be inhabited by several bunyaviruses, as P. cactorum and Phytophthora condilina have been found to host three and 13 putative species, respectively [10,13]. A virus with affinities to Totiviridae has been found to be the most common virus in P. cactorum [10,14], and two related viruses have also been found infecting P. condilina [13]. A member of the proposed genus “Ustivirus” has also been reported from a single isolate of P. cactorum [10]. Viruses infecting the potato late blight pathogen Phytophthora infestans include viruses resembling Narnaviridae [15] and the proposed family “Fusagraviridae” [16], as well as two novel unclassified viruses [17,18]. A novel virus with a resemblance to members of Ghabrivirales was also recently reported from Phytophthora pluvialis [19]. In addition, multiple novel viruses were recorded recently in Phytophthora castaneae, including viruses with phylogenetic affinities to families Endornaviridae, Narnaviridae, Totiviridae, Megabirnaviridae, and “Fusagraviridae” in addition to members of Bunyavirales [20]. A giant virus has also been found integrated in the Phytophthora parasitica genome [21]. In the first population-level study on Phytophthora viruses [10], virus prevalence in a P. cactorum population infecting garden strawberry was found to be very high. The most common viruses were toti-like and endornaviruses occurring in 94% and 57% of isolates, respectively, whereas bunya-like and usti-like viruses occurred much more rarely (4.5% and 1.1% of isolates, respectively). Viral populations infecting isolates from birch and strawberry were also suggested to be different [10].
Although mycoviral infections are most often latent, some mycoviruses can cause phenotypic debilitation in their plant-pathogenic host fungi, and those have been widely studied in the search of potential biocontrol agents [22,23,24,25]. The most successful virocontrol application has been the artificial introduction of hypovirulence-causing Cryphonectria hypovirus 1 (CHV1) to the host fungus Cryphonectria parasitica to control Chestnut blight disease in Europe [26]. The effects of mycoviruses can also be seen as alterations in gene expression. For example, host debilitation inducing Heterobasidion partitivirus 13 and hypovirulence causing Rosellinia necatrix megabirnavirus 1 significantly affected the transcription of 683 and 1160 genes in their hosts, respectively [23,27]. Transcriptome analyses of Cryphonectria hypoviruses, Fusarium graminearum, and Botryosphaeria dothidea viruses as well as Sclerotinia sclerotiorum debilitation-associated RNA virus revealed that the expression levels of specific genes varied between virus-free and virus-infected isolates, but also between phylogenetically different viruses as well as among virus and host strains [28,29,30,31,32]. Despite the importance of oomycete pathogens, very little is known about viral effects on these hosts. However, coinfection of two endornaviruses in Phytophthora sp. from asparagus has been shown to enhance sporangia production in at least some host isolates, inhibit mycelial growth, and potentially modify fungicide sensitivities of the host [11]. Phytophthora infestans RNA virus 2 (PiRV-2), a novel unclassified virus, has also been shown to stimulate sporangia production in P. infestans [33]. It also induced the upregulation of the host ribosomal and histone protein genes, as well as genes for flagellum-related proteins, proteins with epidermal growth factor (EGF)-like conserved site, and genes involved in the glycolytic process in P. infestans [33].
The order Bunyavirales includes viruses with linear, single-stranded, negative-sense, or ambisense RNA genomes classified into 14 families, four subfamilies, 60 genera, and 496 species [34]. Their genomes generally comprise three unique molecules designated L (large), M (medium), and S (small), with sizes of 11–19 kb. The large segment generally codes for an RNA-dependent RNA polymerase (RdRP), whereas the M segment includes genes for two external glycoproteins, and the S-segment encodes a putative nucleocapsid and silencing suppressors [35]. The diversity of bunyaviruses has been increasing due to metagenomic studies [36], and currently, classified species of Bunyavirales are found in plants, invertebrate, and vertebrate hosts. They are also transmitted by arthropod and mammalian vectors [37]. Viruses with genomic similarities to bunyaviruses have more recently also been described from fungi and oomycetes. Fungal species hosting these unclassified bunya-like viruses include several potential phytopathogens [38,39,40,41,42] as well as plant endophytes [43]. Further hosts include the shiitake mushroom (Lentinula edodes) [44] and a marine fungus, Penicillium roseopurpureum [45]. In recent years, bunyaviruses have also been detected in Oomycetes. Pythium polare bunya-like RNA virus 1 (PpBRV1) was reported infecting Globisporangium (syn. Pythium) polare [46], and eight novel (−)ssRNA viruses resembling members of Bunyavirales were recently described in Halophytophthora [47]. Bunya-like viruses were also reported from Plasmopara viticola lesions in grapevine [48]. Furthermore, 13 bunya-like viruses have been detected in P. condilina [13], three in P. cactorum [10], and one in P. castaneae [20]. No direct effects of bunyavirus infection on the growth or virulence of their fungal or oomycete hosts have been recorded. However, Wang et al. [49] detected Macrophomina phaseolina mycobunyaviruses 1–4 (MpMBV1–4) only in hypovirulent fungal isolates.
Fungal viruses usually do not have extracellular infective particles, but they replicate in the cytoplasm of their hosts and transmit intracellularly during cell division and hyphal anastomosis (horizontal transmission) as well as by sexual and asexual spores (vertical transmission) [34,50]. They are usually readily transmitted in vitro between compatible and even incompatible fungal isolates in dual cultures. Successful transmission of an oomycete virus, the unclassified Phytophthora infestans RNA virus 2 (PiRV-2), through a hyphal anastomosis between isogenic hosts has also been reported by Cai et al. [33]. However, the virus did not transfer between different host genotypes, indicating possible hyphal incompatibility between isolates. Nevertheless, PiRV-2 was transmitted to individual zoospores with 100% frequency [34]. In the present study, we aimed to investigate virus effects on P. cactorum to identify oomycetal viruses with biocontrol potential. For this purpose, we tested methods for curing and transferring Phytophthora viruses. The effects of bunyaviruses were examined by comparing the successfully obtained isogenic virus-infected and virus-free isolates that were investigated for their growth rate, sporulation, gene expression, protein abundance, and pathogenicity on two host species (strawberry and birch).

2. Materials and Methods

2.1. Phytophthora Isolates

P. cactorum isolates used in this study were obtained from the isolate collection of Natural Resources Institute Finland (Table 1). They had been isolated from strawberry plants showing crown rot symptoms in three farm locations. The viruses hosted by these isolates have been reported previously [10].

2.2. Curing Virus Infections

Water availability-related matrix stress was induced by a series of different polyethylene glycol (PEG) 8000 (VWR LifeScience) concentrations added to potato dextrose broth (PDB; DifcoTM) growth medium (modified from [51]). Water potential gradients ranging from -1 MPa to -15 MPa were created. The amount of PEG 8000 in a gram of cultivation broth was calculated based on the Michel equation: Ψ (water potential) = 1.29 [PEG]2T–140[PEG]2–4 [PEG], and the value was adjusted to the culture temperature of 20 °C [52]. An agar plug with actively growing mycelium was placed in 50 mL of autoclaved PDB media containing a targeted concentration of PEG 8000. Three isolates (PhF9, PhF66, PhF79; Table 1) hosting endornaviruses (+ssRNA), toti-like viruses (dsRNA), bunyaviruses (-ssRNA), and an usti-like virus (dsRNA) in two replicates and water potentials ranging from −1 MPa to −15 MPa were included in the test. After three weeks, pieces of newly growing mycelia from the liquid media were transferred to potato dextrose agar (PDA; DifcoTM). After one week, the isolates were transferred on modified orange serum (MOS) agar (HiMedia Laboratories) covered with cellophane membrane. After two weeks of growth, mycelia were harvested for RNA extraction. RNA was extracted and reverse transcribed according to Poimala et al. [10], and the resulting cDNAs were screened with PCR for the presence of viruses using specific primers [10]. The disappearance of the viruses was further confirmed by RNA seq, where RNA was extracted with a Spectrum plant total RNA kit (Sigma–Aldrich, Saint Louis, MO, USA) and sent to a sequencing facility (Macrogen, Seoul, Rep. of Korea) for library preparation and RNA-seq, as previously reported [10]. The isolate cured from viruses was designated asPhF66–.
The presence and absence of the bunyaviruses in PhF66 and PhF66–, respectively, was also re-checked before morphological, sporulation, and proteomic analyses. Here, RNAzol (Sigma-Aldrich) was used for RNA extraction and a High-capacity cDNA kit (Thermo Fisher) for reverse transcription, as in Raco et al. [20].

2.3. Virus Transfer Experiment

Isolates PhF38, PhF66, PhF79, and PhF101 (Table 1) were used as donor strains and plated as dual cultures with the virus-free recipient PhF66– with six replicates on malt agar plates to enable the formation of hyphal contacts and subsequent viral transmission. Mycelial samples were taken from the recipient side of the culture after 21 and 64 days and analyzed for the presence of the corresponding viruses using RNA extraction and RT-PCR with virus-specific primers using the original virus-containing PhF66 as a positive control [10].

2.4. Growth Tests

The phenotypes of PhF66 and PhF66– were compared by testing their growth rates on 2% MEA plates at 20 °C. The tests were conducted using 22 replicates of each isolate. Each isolate was inoculated in the center of the plate as a mycelial plug with a diameter of 5 mm. Mycelial areas were marked on the plates two and six days after growth initiation, after which the plates were scanned. The surface area covered by the fungal mycelium was determined from the images using ImageJ program. Radial growth rates were calculated in Excel. Furthermore, 12 apple fruits (‘Golden Delicious’) per isolate were inoculated by drilling a 10 mm hole to the surface with a Ø 10 mm cork borer and inserting a mycelial plug containing mycelia (Ø 5 mm) from a 13-day-old culture to the bottom. After six days at +22 °C, the areas of the brown lesions emerging on the apple surfaces were measured. The differences in growth between the isolates were determined by one-way ANOVA (growths on MEA) or independent samples Mann Whitney U-test (lesions in apple) in IBM SPSS Statistics v27.

2.5. Sporulation Tests

The isolates PhF66 and PhF66– were grown on clarified 11 mL vegetable juice agar [VA; 100 mL/L vegetable juice DM Bio Gemüse Saft (DM-Drogerie Markt, Karlsruhe, Germany), 3 g/L of CaCO3, 18 g/l of agar (VWR LifeScience)] in 90 mm diameter Petri dishes at 20 °C in the dark. To obtain zoospores, three experiments were conducted. In experiment 1, three replicates per isolate were used. For each of them, a mycelial plug containing mycelia (Ø 5 mm) was placed in the center of a Petri dish containing 11 mL clarified VA. After four and five weeks, cultures were flooded with 12 mL sterile demineralized water for 24 to 48 h under daylight at 20 °C. Water was decanted and replaced two to three times per day. For experiments 2 and 3, two 15 × 15 mm agar plugs from three-, five-, and six-day-old cultures (four plates per isolate; two plugs per plate) were cut from the same area for each replicate (one cut closer to the edge of a Petri dish, and the other closer to the agar plug) were placed in an empty 90 mm diameter Petri dish and flooded with 12 mL of sterile demineralized water. In experiment 2, the water was replaced after 12 h, 16 h, 20 h, and 22 h with 10% pond water (100 mL/L non-sterile pond water, 900 mL distilled autoclaved water). In experiment 3, the water was replaced with sterile demineralized water at 3.5 h, with Volvic® mineral water (Danone, Clermont-Ferrand, France) at 7 h and 21 h, and with Volvic® mineral water amended with approximately 1 mL of 2 days old non-sterile soil extract (250 mL of soil taken under asymptomatic Juglans regia L. trees flooded with 2 L of water) at 24 h. After the incubation period under daylight at 20 °C, zoospore formation and release were stimulated by cold shock (60 min in +4 °C). Zoospore suspension was prepared, and spore counts were determined, as in Harris & Webber [53]. A longer incubation of 120 min at 7 °C prior to the 75 min incubation at 20 °C was also tested for stimulating zoospore release. The agar plates (experiment 1), and agar plugs (experiments 2 and 3) were inspected under a microscope to follow sporangia formation before each change of water. Zoospore counts were determined in a Bürker counting chamber. To evaluate sporulation levels, four 0.5 cm × 0.5 cm squares were cut from the four compass points on a Petri dish with five-day-old mycelia from the area of 0.5 cm to 1.0 cm from the center of the agar plug. Sporangia formation was stimulated as in the experiment number 3. The number of sporangia produced by PhF66 and PhF66– was estimated by counting all sporangia, including freshly forming, already formed (mature), and open sporangia in these four agar squares on one plate per isolate. Statistical differences were obtained by two sample t-test in R (v 3.6.3).

2.6. Morphology

The dimensions of 50 randomly selected sporangia and 50 gametangia (oogonia and antheridia) were measured from the isolates PhF66 and PhF66–. Microscopic examinations and measurements were performed as in Jung et al. [54,55], except that the sporangia for measurements were produced as in the sporulation experiment 3. Characteristics of mature gametangia (antheridia and oogonia) were examined on clarified VA after 28–35 days at 20 °C in the dark. Abortion rates were estimated from three plates per isolate by counting 100 oogonia in two 15 × 15 mm squares, one cut from the area closer to the edge, and the other from the area closer to the center of the plate. Microscopic measurements of all structures at ×400 were made using an optical microscope (Motic© BA410E), camera (Moticam 5 + 5.0 MP), and camera software (Motic© Images Plus 3.0). The photographs were taken using a compound microscope (Zeiss Axioimager.Z2, Carl Zeiss AG, Oberkochen, Germany), a digital camera (Zeiss Axiocam ICc5), and biometric software (Zeiss ZEN). Statistical differences were obtained by Bonferroni corrected two sample t-test in R (v 3.6.3).

2.7. RNA Extraction and RNA-Seq

Prior to total RNA extraction, the isolates were grown on MOS agar plates on cellophane membranes. The mycelia were harvested after two weeks, freeze-dried, and ground in liquid nitrogen, and total RNA was extracted using Spectrum Plant total RNA kit (Merck) from three technical replicates of PhF66 and PhF66– (i.e., mycelia originating from one plate). These six samples were considered biological replicate 1. Furthermore, RNA from two more biological replicates of the isogenic strains (in two technical repeats) were extracted separately in the same way, resulting in four additional RNA-seq samples. The RNA of each sample was quantified with a NanoDrop 2000 Spectrophotometer (Thermo scientific, Waltham, MA, USA). The samples were sent to a sequencing facility (Novogene UK Company Ltd., London, UK), where RNA integrity was assessed, eukaryotic strand-specific transcriptome libraries were built, and sequencing was performed using the Illumina NovaSeq6000 system, which generates stranded paired-end sequences.

2.8. Bioinformatics

The RNA-Seq data analysis was performed in Chipster software [56]. Phred score was >35 along all reads in every sample, so no trimming was needed before read mapping. Reads were mapped to P. cactorum reference genome (GCA_010194725.1) with HISAT2. Reads were counted using HTSeq, and differential expression calls were made with edgeR and DESeq2 in Chipster v4 software. The raw reads are available in GenBank under BioProject PRJNA804427.

2.9. Proteomics

For proteome analysis, isolates were grown in liquid cultures (autoclaved mixture of 100 mL/L of vegetable juice premixed with 2 g/L of CaCO3 and 900 mL/L distilled H20) for nine days, thoroughly washed in Tris-buffered saline (TBS) (150 mM NaCl, 50 mM Tris-HCl, pH 7.6) to remove any residues of the cultivation medium, then frozen in liquid nitrogen and lyophilized. The proteome analysis was performed with liquid chromatography-mass spectrometry as previously described [57]. In brief, lyophilized samples were extracted with tert-butyl methyl ether/methanol mixture, proteins were solubilized, digested with trypsin, and analyzed by nanoflow reverse-phase LC-MS using a 15 cm C18 Zorbax column (Agilent, Santa Clara, CA, USA), a Dionex Ultimate 3000 RSLC nano-UPLC system, and the Orbitrap Fusion Lumos Tribrid Mass Spectrometer (Thermo Fisher, Waltham, MA, USA). The measured spectra were recalibrated and compared against the databases of P. cactorum [58] and common contaminants using Proteome Discoverer 2.5 (Thermo Fisher) with Sequest HT and MS Amanda 2.0 [59] algorithms. The quantitative differences were determined by Minora, employing precursor ion quantification followed by normalization and calculation of relative peptide/protein abundances. The results were evaluated by a background-based t-test, and the resulting p-values were adjusted using the Benjamini-Hochberg method. Only proteins with at least two unique peptides were considered for the quantitative analysis. The mass spectrometry proteomics data were deposited at the ProteomeXchange Consortium via the PRIDE [60] partner repository with the data set identifier PXD033832.

2.10. Pathogenicity

In order to test for virulence of PhF66 and PhF66– to strawberry, two inoculation tests were set up in a greenhouse in October-December 2020 (strawberry test I). Runner plugs of greenhouse-maintained strawberry plants of varieties Glima, Jonsok, Senga Sengana, Zefyr, and Honeoye were rooted in rockwool grow cubes (2 × 2 × 3 cm; Grodan, Roermond, the Netherlands) two weeks prior to inoculation. The well-rooted plug plants were then randomized and spaced into larger rockwool plates fitted in plastic containers, which were subsequently kept on a table in a greenhouse bench under +23 °C and long day (19 h) conditions. Six plants of each variety per each inoculation treatment were wounded in the stem (crown) base with a sterile blade tip, and a malt agar plug containing mycelium of PhF66, PhF66– or sterile agar plugs containing no hyphae (mock-inoculated control plants) was placed on the wound. The inoculum was covered with moist cotton and sealed with parafilm. The cotton was removed after five days. The experiment was monitored for 30 days, and repeated twice, two weeks apart. Strawberry test II was set up in June 2021 with strawberry plants (Malling Centenary) planted in the previous summer and growing outdoors in plastic growth bags filled with peat (Kekkilä Growth Sack Airboost 1 m). The plants were inoculated on 28 May 2021, as described above, with an agar plug containing PhF66 (four plants), PhF66– (four plants), PhF17/19 (six plants) –another P. cactorum strain isolated from strawberry [10]– or no hyphae (mock-inoculated control; four plants). Leaves were counted and scored healthy/wilting/dead once a week for three weeks. Furthermore, one-year-old silver birch (Betula pendula) seedlings were inoculated with isolates PhF66 (15 seedlings), PhF66– (15 seedlings), Ph415 –another P. cactorum strain isolated from birch (Table 1; [61])–(13 seedlings), and agar-control (10 seedlings) in June 2021. Birch stem inoculations were performed similarly to that of strawberries, by wounding the bark and placing an malt agar plug containing hyphae onto the wound [62]. The longitudinal lengths of the resulting lesions were measured at 6, 13, and 20 days post infection. Differences in lesion sizes between isolates were analysed by one-way ANOVA in IBM SPSS Statistics.

3. Results

3.1. Curing Virus Infections

All isolates grew in all water matrix potentials applied in the experiment, but their growth was significantly reduced at −13 MPa and −15 MPa compared to growth rates at −11 MPa. The two replicates of each isolate (PhF9, PhF66, and PhF79) recovered from water potentials −13 MPa and −15 MPa were subjected to RNA extraction, reverse transcription, and virus-specific PCR. Amplification products of alphaendornaviruses PcAEV1, PcAEV2, and PcAEV3 were obtained from all replicates of the isolate hosting them (PhF9). The toti-like PcRV1 was also detected from all replicates of isolates PhF9 and PhF79. Phytophthora cactorum usti-like virus (PcUV1) as well as the bunyavirus PcBV2 were detected in all PhF79 replicates. However, the bunyavirus PcBV1 was not detected in PhF79, and both PcBV1 and PcBV2 were cured from all four analyzed replicates of PhF66. The result was confirmed by RNA-seq analysis of isolate PhF66–, where no reads of virus sequences were obtained.

3.2. Transfer Tests

Mycelial samples taken from the recipient (the virus-free PhF66–) side of the dual cultures with virus hosting strains (PhF38, PhF66, PhF79, and PhF101) were analyzed for the presence of transmitted viruses using RT-PCR with virus-specific primers [10]. Viruses PcEV1, PcAEV2, PcAEV3, PcBV1, PcBV2, PcBV3, PcRV1, and PcUV1 were screened according to the viruses in the donor, but no amplification was observed at 21 or 64 days. Therefore, it was concluded that no successful virus transmissions had occurred.

3.3. Viral Effects on Host Growth, Sporulation and Morphological Characteristics

To test the phenotypic effects of bunya-like viruses in laboratory conditions, we conducted growth experiments using strains PhF66 and PhF66–, hosting two bunyaviruses and no viruses, respectively. The presence of these viruses significantly lowered the growth rate of the host isolate PhF66 in both growth substrates. There was a significant effect of the virus status on the growth rate of PhF66. The radial growth rate of the bunyavirus-containing PhF66 on MEA was 2.3 mm/day (range 2.20–2.51 mm/day) and that of PhF66– was 3.1 mm/day (range 2.89–3.24 mm/day), which differed significantly from each other (F = 936, p < 0.001, N = 44).The mean lesion areas on apples were 10.0 cm2 (range 1.85–18.18 cm2) and 15.8 cm2 (range 1.97–31.14 cm2), respectively, which also differed significantly (U = 149, p < 0.05, N = 44)
To examine the effects of the two bunya-like viruses on sporangia formation and zoospore release, three experiments using different settings were conducted. Sporangia were produced by all three methods, but no reliable zoospore counts were obtained. In experiment 1, the average number of encysted zoospores in the 12 uL examined for each replicate in Bürker chamber was 2.66 for PhF66 and 4.2 for PhF66–, which was considered too low for a reliable count. In experiments 2 and 3, empty sporangia and a few motile zoospores (Figure 1a) were observed under a microscope, but later on no reliable counts of encysted zoospores were obtained. At the time of the expected zoospore release, some sporangia germinated directly by the formation of a germination tube instead of producing zoospores (Figure 1b). In addition, in many cases, there was no evidence of zoospore differentiation inside the sporangia, and cytoplasm would often be released out of them instead of biologically viable zoospores (Figure 1c). In experiment 2, the formation of young sporangia was recorded in both isolates and all replicates at 16 h after the first flooding.
Differences in sporangia production and morphological characteristics were observed between isolates PhF66 and PhF66–. PhF66 produced significantly less (p < 0.01) and smaller (p < 0.0001) sporangia compared to PhF66– (Table 2). No differences between the isolates were found in the pedicel size, oogonia size, oospore diameter, thickness of oospore walls, or antheridia size. Oospore abortion rates were relatively high in both variants but significantly higher (p < 0.001) in the virus-free PhF66– (Table 2).

3.4. Viral Effects on the Host Gene Expression and Protein Abundance

The overall alignment rate of the raw RNA-seq reads to the P. cactorum reference genome was 94.81–95.27% between the samples. While summarizing mapped reads into a gene level count, 51.0–54.4% of reads were not counted in the samples (mostly due to non-unique alignments and less than complete annotation of the reference genome). All three biological replicates of PhF66 differentiated significantly from those of PhF66– in the principal component analysis and heatmap (Supplementary File S1). A batch effect was also found between the first biological replicate (six samples) and biological replicates 2 and 3 (four samples) (Supplementary File S1), which was taken into account as an additional factor in the differential expression analysis. The number of differentially expressed genes (DEGs) having p-values < 0.05 with both egdeR and DESeq2 was 1824. Out of the 1824 DEGs, 120 had cutoff values greater than 1.5 (Supplementary Table S1), and in the case of 33 genes, the cutoff value was greater than 2 (representing over 4× fold change) (Table 3). Ten of the 33 DEGs represented hypothetical proteins, but a conserved domain was found in 23 DEGs. Overall, 10 genes were upregulated in the virus-containing isolate, three of which were elicitin genes (Table 3). Notably, no transcripts of the gene coding for hypothetical protein GQ600_18332 were obtained from the virus-free isolate PhF66–, but it was abundantly transcribed in PhF66 (Table 3). The amino acid sequence of protein GQ600_18332 had 43.8% identity with the DDE_Tnp_1_7 domain containing protein of Phytophthora palmivora in Uniprot BLAST. The domain is found in the PiggyBac transposable element-derived proteins (PGBDs) as well as in some uncharacterized proteins, but its function is not known [63]. Among the DEGs with lower cut-offs (1.5–2), upregulated genes also included, e.g., a CAP (cysteine-rich secretory proteins, antigen 5, and pathogenesis-related 1) domain that likely functions in sequestering sterols from plants [64,65], a C2H2 zinc finger domain that has putative roles in oospore development and virulence [66], and a “necrosis inducing protein” which refers to NPP1-like proteins in Oomycetes capable of inducing ROS and HR-like cell death in the host [67] (Supplementary Table S1).
Out of the 23 genes that were downregulated over 4× fold in the virus-containing isolate, four included a WD40 repeat-containing domain (Table 3). WD40 domain-containing proteins are found in all eukaryotes, and they function as coordinators of protein–protein interactions [68]. The ankyrin-repeat domain (found among up- and downregulated DEGs) has a similar function [69]. Among the DEGs with cutoffs 1.5–2, downregulated genes included, e.g., two additional WD40 repeat domains and three genes with Leucine-rich repeat (whose function is largely unknown in Oomycetes but reported to be involved in zoospore flagellum development in P. sojae [70]).
In the proteomic analysis, 1392 proteins were detected, out of which 53 differed significantly (p < 0.05) in their abundance between PhF66 and PhF66–. The differentially abundant proteins included, e.g., two elicitins and various enzymes (Table 4). When comparing the data from transcriptomic and proteomic analyses, only one of the 33 DEGs with cut-offs >2 were found among the 53 differentially abundant proteins–namely an elicitin. Two additional differentially expressed genes were identified among the detected non-differentally abundant proteins. When searching for the 53 differentially abundant proteins in the RNA-seq data, only six of them were found among the 1824 differentially expressed genes (p < 0.05) (Table 4).

3.5. Viral Effects on Host Pathogenicity

The virulence of isolates PhF66 and PhF66– to strawberry was tested in two occasions. In the strawberry test I which included five strawberry cultivars and two repetitions, no typical symptoms of crown rot appeared during the 30-day experiment. Cross-sections of crown stems of the strawberry plants were inspected, and re-isolation of the pathogen was attempted at 30 dpi. However, no lesions or browning were found in the internal tissues, and pathogen isolations were unsuccessful. In the strawberry test II, the first wilting of leaves was observed on two plants at 4 dpi, on plants inoculated with PhF17/19 (positive control), and the first dead plant was observed at 12 dpi. The last inventory was made at 20 dpi, when all plants inoculated with PhF17/19 were dying. No symptoms were found on plants inoculated with PhF66 or PhF66– (Figure 2a,b). Re-isolations were attempted at 20 dpi from all inoculations, but they were successful only from plants infected with PhF17/19 (four re-isolations). The results are shown in Table 5.
In the pathogenicity test with silver birch, lesions in all inoculated plants were observed at 6 dpi (Figure 2c). However, a very small increase in lesion length was measured at 13 dpi. No lesion growth was recorded after 13 dpi, indicating the subsequent healing of lesions most likely due to very high temperatures (over 30 °C) for several days at the time of the test, which may have slowed down the pathogen growth and allowed the seedlings to suppress its advancement in the stem tissues. Re-isolations were attempted at the time of last inventory at 20 dpi, but with no success. ANOVA was performed with 13 dpi measurements, which returned no significant differences in lesion sizes induced by the three isolates (F = 0.077, p = 0.926, N = 41, Figure 3). No lesion development was detected among the mock-inoculated control seedlings.

4. Discussion

Bunya-like viruses have been described from the oomycetes P. condilina [13], P. castaneae [20], Halophytopththora spp. [47], Globisporangium polare [46], as well as many fungal hosts, but to our knowledge there are no previous reports on the direct effects of bunyavirus infections on fungal or oomycete hosts. In the present study, infection by two bunyaviruses (PcBV1 & PcBV2) reduced the growth rate of the host isolate (PhF66) on MEA plates as well as in apple tissue. Two endornaviruses have also been shown to inhibit mycelial growth in Phytophthora sp. from asparagus, but in contrast to the results of the current study, sporangia production was enhanced [11].
Both transcriptomic and proteomic analyses indicated that certain elicitins were differentially expressed in the virus-infected and virus-free isolate. The expressions of three elicitins were strongly upregulated at the RNA level, and one of them was also more abundant at the protein level in the virus-containing isolate (Table 3 and Table 4). Elicitins are structurally conserved proteins excreted by Oomycetes, mainly by species of Phytophthora and Pythium, during host plant interaction. They are apoplastic effectors that are utilized to suppress induced immunity responses in the host plant [71,72,73]. While elicitins promote oomycete virulence, in some pathosystems, they can in turn become recognized by the host defense system and become associated with reduced pathogenicity [74,75]. The translation initiation of a gene putatively coding for a PiggyBac transposable element-derived protein in the virus-containing isolate may indicate viral effect on transposon activity in the host, although no proteins including this domain were found among the proteins differing in their abundance. The downregulated DEGs in the virus-containing isolate included many genes that function as mediators of protein–protein interactions in multi-protein complexes (WD40 and ankyrin repeat domains). While WD40 domain-containing proteins have been identified as negative regulators of Arabidopsis resistance to Phytophthora [76,77], it is also found in, e.g., cyclophilin-encoding genes in the Phytophthora species, which potentially have various roles in plant infection [78]. The WD40 domain is also found in the avirulence genes of P. infestans, P. sojae, and P. ramorum, coding for cytoplasmic effectors that trigger an immune response in the host [79,80,81].
Relatively few of the DEGs were found to be differentially abundant as proteins and vice versa. While RNA sequencing enables the profiling of slight changes in the transcript expression between two conditions, it does not capture the involvement of various post-transcriptional, post-translational, and degradation pathways that influence the relative quantities of expressed protein [82]. Thus, RNA transcript and protein abundance are often only weakly correlated [83]. Furthermore, the mRNAs (differentially) expressed at low levels may be below the detection limit at the protein level. It should be noted that the isolates were grown on V8 growth media instead of MOS prior to the proteomic analyses, which likely resulted in some differences in the gene and protein expressions.
In the pathogenicity tests conducted with the bunyavirus-containing and virus-free isolates (PhF66 and PhF66–), both isolates failed to cause symptoms in strawberry. However, lesions were induced on silver birch indicating the isolate may belong to the population of P. cactorum preferring birch as a host. However, we did not observe any differences between the virus-free and virus-containing isolates in pathogenicity to birch. Thus, understanding the roles and functions of the elicitins upregulated by these bunyaviruses requires further investigation.
The PhF66 isolate containing the bunyaviruses also produced less and smaller sporangia. However, no effective zoospore release was documented. Stimulation of sporangia production was also reported in P. infestans by the unclassified virus PiRV-2 [33]. The present study also found relatively high oospore abortion rates in both isolates (PhF66 & PhF66–), which was significantly higher in the virus-free isolate. In an evolutionary analysis of European P. cactorum populations, Pánek et al. [84] inferred Finnish P. cactorum isolates (representing woody species pathotype) to possibly be of hybrid origin. The isolates were also characterized by high oogonial abortion rates, suggesting that the putative hybrid status could be the reason for the relatively high oospore abortion rates detected in the present study. The higher oospore abortion rate in the virus-free isolate may also be related to stress caused by the virus-curing treatment in low water potential. Adverse conditions such as the long deep-freeze storage or previous exposure to fungicides in the field can decreaseoospore viability (T. Jung, pers. comm.).
In this study, we found that the most prevalent virus infections (endorna- and toti-like viruses) in P. cactorum are highly stable and resist curing attempts by osmotic stress, but the rarer infections of bunyaviruses seem less stable and more easily curable. In general, the curing and transmission of viruses in vitro has been challenging with the Phytophthora species [33,85]. In the present study, to avoid mutations in the host caused by chemical treatments such as Ribavirin, we tested the suitability of PEG-induced stress on water availability for curing viruses from Phytophthora. We found that the P. cactorum isolates were very persistent during low water availability conditions and were able to grow in and were recoverable from all tested water potentials, down to −15 MPa. The method was originally developed and proved successful for the fungus Pseudogymnoascus destructans, whose growth ceased at −5 MPa, and an inhabiting partitivirus was cured already at −2 MPa [51]. In our study, the more nutrient-rich medium (potato dextrose) could have favored the growth of Phytophthora isolates and thus explain its viability in the very low water potentials during this study. All three alphaendornaviruses (+ssRNA; PcAEV1-3), a toti-like virus (dsRNA; PcRV1), and an ustivirus (dsRNA; PcUV1) remained in the host isolates during the treatment. However, bunyaviruses (-ssRNA; PcBV1, PcBV2) were shown to be curable with this method. Previously, the unclassified dsRNA virus Phytophthora infestans RNA virus 2 (PiRV-2) and a member of “Fusagraviridae”, Phytophthora infestans RNA virus 3 (PiRV-3), were cured successfully after four and five hyphal tip transfers on Ribavirin-containing plates, respectively [16,33]. Curing was reported unsuccessful for a narnavirus (PiRV4, [15]) and the unclassified Phytophthora infestans RNA virus 1 (PiRV1, [85]). In the present study, the duration of the treatment was limited to three weeks. The usefulness of additional combinations of water potentials, lower nutrient media, and longer test duration for curing additional viruses from Phytophthora may be subjects of further study. In conclusion, PEG-induced water stress is a potential method for curing Phytophthora viruses, but efficacy varies between viral species.
Despite viruses being highly common in isolates of P. cactorum [10], we found that they are not readily transmissible by hyphal contacts in laboratory conditions, as the virus transfer between P. cactorum isolates in dual cultures was not successful. Virus transmission by hyphal anastomosis has been attempted but also proved relatively challenging for P. infestans [85]. However, Cai et al. [33] reported a successful reintroduction of PiRV-2 into the same isolate after two weeks on joint rye agar plates. In the present study, the dual cultures of different strains were maintained for 64 days on MEA, and no virus transfer was observed even between isogenic PhF66 and PhF66–. Overall, it seems that the viral transfer via mycelial contact is not a frequent event, even between isogenic Phytophthora isolates in laboratory conditions. At a population level, P. cactorum was recently shown to host multiple viruses with an extremely high prevalence (99% of isolates hosted viruses), indicating a long co-evolution likely resulting in neutral or positive symbiotic relationships [10]. Furthermore, the pathotypes infecting strawberry seem to share the same viruses globally, suggesting the transport of P. cactorum along with host plants between continents [10]. However, the presence of the same viruses in multiple host isolates can also result from a higher vertical transmission potential between host genotypes in the field compared to laboratory conditions, and it should also be noted that species of Phytophthora are prone to hybridization [86], which could provide an additional route for virus transmission.
In summary, the present study showed that bunyaviruses, but not other tested viruses, may be cured from P. cactorum by PEG-induced water stress, although transmission between different mycelia in laboratory conditions seems to be inefficient. Bunyaviruses were shown to cause phenotypic effects on the growth rate of P. cactorum on artificial media and in apple fruits, but not on the pathogenicity. Bunyaviruses also induced the upregulation of elicitin genes, one of them at both transcript and protein levels, in the virus-containing isolate. Upregulation of other apoplastic effectors (CAP-family, necrosis-induced proteins) was also observed at the transcriptional level, indicating a putative viral effect on host pathogenesis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/v14122596/s1, Supplementary File S1: Principal component analysis (PCA) and heatmap of gene counts after alignment of 5 + 5 (of PhF66 and PhF66–, respectively) samples of RNA-seq reads to the reference genome. Supplementary Table S1: DEGs with cutoffs >1.5.

Author Contributions

Conceptualization, E.J.V., J.H., A.P., M.R.; methodology, A.P., T.H., P.P., E.J.V., M.R., M.Č.; formal analysis, A.P., M.R., M.Č.; writing—original draft preparation, A.P.; writing—A.P., E.J.V., J.H., T.H., M.R., M.Č. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by European Union’s Horizon 2020 Research and Innovation Programme under Grant Agreement No. 773567. Funding was also received from the European Regional Development Fund, project “Phytophthora Research Centre,” Reg. No. CZ.02.1.01/0.0/0.0/15_003/0000453 and Specific University Research Fund of the Faculty of Forestry and Wood Technology, Mendel University in Brno LDF_VP_2020017.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The RNA-seq raw reads are available in GenBank under BioProject PRJNA804427, and the mass spectrometry proteomics data are available at the ProteomeXchange Central with the data set identifier PXD033832.

Acknowledgments

The authors thank the technical staff at Luke Viikki molecular laboratory for excellent molecular work, and the staff at Luke Piikkiö experimental station for the maintenance of the greenhouse trial. Thomas Jung, Martin S. Mullett, Zoltan Nagy, and Ivan Milenković (Mendel University, Brno), and Matěj Pánek (Crop Research Institute, Prague) are thanked for helpful discussion related to sporulation experiments and morphological features. Thomas Jung is also thanked for photographing P. cactorum morphological structures with the Zeiss compound microscope.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Erwin, D.C.; Ribeiro, O.K. Phytophthora Diseases Worldwide; American Phytopathological Society (APS Press): Saint Paul, MN, USA, 1996. [Google Scholar]
  2. Nellist, C.F.; Vickerstaff, R.J.; Sobczyk, M.K.; Marina-Montes, C.; Wilson, F.M.; Simpson, D.W.; Whitehouse, A.B.; Harrison, R.J. Quantitative trait loci controlling Phytophthora cactorum resistance in the cultivated octoploid strawberry (Fragaria × ananassa). Hortic. Res. 2019, 6, 60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Eikemo, H.; Stensvand, A.; Davik, J.; Tronsmo, A.M. Resistance to crown rot (Phytophthora cactorum) in strawberry cultivars and in offspring from crosses between cultivars differing in susceptibility to the disease. Ann. Appl. Biol. 2003, 142, 83–89. [Google Scholar] [CrossRef]
  4. Eikemo, H.; Stensvand, A.; Tronsmo, A.M. Induced resistance as a possible means to control diseases of strawberry caused by Phytophthora spp. Plant Dis. 2003, 87, 345–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Bourret, T.B.; Fajardo, S.N.; Engert, C.P.; Rizzo, D.M. A barcode-based phylogenetic characterization of Phytophthora cactorum identifies two cosmopolitan lineages with distinct host affinities and the first report of Phytophthora pseudotsugae in California. J. Fungi 2022, 8, 303. [Google Scholar] [CrossRef] [PubMed]
  6. Bhat, R.G.; Colowit, P.M.; Tai, T.H.; Aradhya, T.H.; Browne, G.T. Genetic and pathogenic variation in Phytophthora cactorum affecting fruit and nut crops in California. Plant Dis. 2006, 90, 161–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Eikemo, H.; Klemsdal, S.S.; Riisberg, I.; Bonants, P.; Stensvand, A.; Tronsmo, A.M. Genetic variation between Phytophthora cactorum isolates differing in their ability to cause crown rot in strawberry. Mycol. Res. 2004, 108, 317–324. [Google Scholar] [CrossRef] [Green Version]
  8. Hantula, J.; Lilja, A.; Parikka, P. Genetic variation and host specificity of Phytophthora cactorum isolated in Europe. Mycol. Res. 1997, 101, 565–572. [Google Scholar] [CrossRef]
  9. Chepsergon, J.; Motaung, T.E.; Bellieny-Rabelo, D.; Moleleki, L.N. Organize, don’t agonize: Strategic success of Phytophthora species. Microorganisms 2020, 8, 917. [Google Scholar] [CrossRef] [PubMed]
  10. Poimala, A.; Parikka, P.; Hantula, J.; Vainio, E.J. Viral diversity in Phytophthora cactorum population infecting strawberry. Env. Microbiol. 2021, 23, 5200–5221. [Google Scholar] [CrossRef] [PubMed]
  11. Uchida, K.; Sakuta, K.; Ito, A.; Takahashi, Y.; Katayama, Y.; Omatsu, T.; Mizutani, T.; Arie, T.; Komatsu, K.; Fukuhara, T.; et al. Two novel endornaviruses co-infecting a Phytophthora pathogen of Asparagus officinalis modulate the developmental stages and fungicide sensitivities of the host oomycete. Front. Microbiol. 2021, 12, 633502. [Google Scholar] [CrossRef] [PubMed]
  12. Hacker, C.V.; Brasier, C.M.; Buck, K.W. A double-stranded RNA from a Phytophthora species is related to the plant endornaviruses and contains a putative UDP glycosyltransferase gene. J. Gen. Virol. 2005, 86, 1561–1570. [Google Scholar] [CrossRef] [PubMed]
  13. Botella, L.; Jung, T. Multiple viral infections detected in Phytophthora condilina by total and small RNA Sequencing. Viruses 2021, 13, 620. [Google Scholar] [CrossRef] [PubMed]
  14. Poimala, A.; Vainio, E.J. Complete genome sequence of a novel toti-like virus from the plant-pathogenic oomycete Phytophthora cactorum. Arch. Virol. 2020, 165, 1679–1682. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, G.; Myers, K.; Fry, W.E.; Hillman, B.I. A member of the virus family Narnaviridae from the plant pathogenic oomycete Phytophthora infestans. Arch. Virol. 2012, 157, 165–169. [Google Scholar] [CrossRef] [PubMed]
  16. Cai, G.; Krychiw, J.F.; Myers, K.; Fry, W.E.; Hillman, B.I. A new virus from the plant pathogenic oomycete Phytophthora infestans with an 8 kb dsRNA genome: The sixth member of a proposed new virus genus. Virology 2013, 435, 341–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Cai, G.; Myers, K.; Hillman, B.I.; Fry, W.E. A novel virus of the late blight pathogen, Phytophthora infestans, with two RNA segments and a supergroup 1 RNA-dependent RNA polymerase. Virology 2009, 392, 5261. [Google Scholar] [CrossRef] [PubMed]
  18. Cai, G.; Myers, K.; Fry, W.E.; Hillman, B.I. Phytophthora infestans RNA virus 2, a novel RNA virus from Phytophthora infestans, does not belong to any known virus group. Arch. Virol. 2018, 164, 567–572. [Google Scholar] [CrossRef] [PubMed]
  19. Xu, Z.; Khalifa, M.E.; Framptom, R.A.; Smith, G.R.; McDougal, R.L.; MacDiarmid, R.M.; Kalamorz, F. Characterization of a novel double-stranded RNA virus from Phytophthora pluvialis in New Zealand. Viruses 2022, 14, 247. [Google Scholar] [CrossRef] [PubMed]
  20. Raco, M.; Vainio, E.J.; Sutela, S.; Eichmeier, A.; Hakalová, E.; Jung, T.; Botella, L. High diversity of novel viruses in the tree pathogen Phytophthora castaneae revealed by high-throughput sequencing of total and small RNA. Front. Micobiol. 2022, 13, 911474. [Google Scholar] [CrossRef] [PubMed]
  21. Hannat, S.; Pontarotti, P.; Colson, P.; Kuhn, M.L.; Galiana, E.; La Scola, B.; Aherfi, S.; Panabières, F. Diverse trajectories drive the expression of a giant virus in the oomycete plant pathogen Phytophthora parasitica. Front. Microbiol. 2021, 12, 662762. [Google Scholar] [CrossRef]
  22. 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]
  23. Vainio, E.J.; Jurvansuu, J.; Hyder, R.; Kashif, M.; Piri, T.; Tuomivirta, T.; Poimala, A.; Xu, P.; Mäkelä, S.; Nitisa, D.; et al. Heterobasidion partitivirus 13 mediates severe growth debilitation and major alterations in the gene expression of a fungal forest pathogen. J. Virol. 2018, 92, e01744-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Grasse, W.; Zipper, R.; Totska, M.; Spring, O. Plasmopara halstedii virus causes hypovirulence in Plasmopara halstedii, the downy mildew pathogen of the sunflower. Fungal Genet. Biol. 2013, 57, 42–47. [Google Scholar] [CrossRef] [PubMed]
  25. Prospero, S.; Botella, L.; Santini, A.; Robin, C. Biological control of emerging forest diseases: How can we move from dreams to reality? For. Ecol. Manag. 2021, 496, 119377. [Google Scholar] [CrossRef]
  26. 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] [PubMed] [Green Version]
  27. Shimizu, T.; Kanematsu, S.; Yaegashi, H. Draft genome sequence and transcriptional analysis of Rosellinia necatrix infected with a virulent mycovirus. Phytopathol. 2018, 108, 1206–1211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. 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] [PubMed]
  29. Allen, T.D.; Dawe, A.L.; Nuss, D.L. Use of cDNA microarrays to monitor transcriptional responses of the chestnut blight fungus Cryphonectria parasitica to infection by virulence-attenuating hypoviruses. Eukaryot. Cell 2003, 2, 1253–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Li, H.; Fu, Y.; Jiang, D.; Li, G.; Ghabrial, S.A.; Yi, X. Down-regulation of Sclerotinia sclerotiorum gene expression in response to infection with Sclerotinia sclerotiorum debilitation-associated RNA virus. Virus Res. 2008, 135, 95–106. [Google Scholar] [CrossRef]
  31. Cho, W.K.; Yu, J.; Lee, K.-M.; Son, M.; Min, K.; Lee, Y.-W.; Kim, K.-H. Genome-wide expression profiling shows transcriptional reprogramming in Fusarium graminearum by Fusarium graminearum virus 1-DK21 infection. BMC Genom. 2012, 13, 173. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, L.; Luo, H.; Hu, W.; Yang, W.; Hong, N.; Wang, G.; Wang, A.; Wang, L. De novo transcriptomic assembly and mRNA expression patterns of Botryosphaeria dothidea infection with mycoviruses chrysovirus 1 (BdCV1) and partitivirus 1 (BdPV1). Virol. J. 2018, 15, 126. [Google Scholar] [CrossRef] [PubMed]
  33. Cai, G.; Fry, W.E.; Hillman, B.I. PiRV-2 stimulates sporulation in Phytophthora infestans. Virus Res. 2019, 271, 197674. [Google Scholar] [CrossRef] [PubMed]
  34. Walker, P.J.; Siddell, S.G.; Lefkowitz, E.J.; Mushegian, A.R.; Adriaenssens, E.M.; Alfenas-Zerbini, P.; Dempsey, D.M.; Dutilh, B.E.; García, M.L.; Hendrickson, R.C.; et al. Recent changes to virus taxonomy ratified by the International Committee on Taxonomy of Viruses. Arch. Virol. 2022, 167, 2429–2440. [Google Scholar] [CrossRef] [PubMed]
  35. Margaria, P.; Bosco, L.; Vallino, M.; Ciuffo, M.; Mautino, G.C.; Tavella, L.; Turina, M. The NSs Protein of tomato spotted wilt virus is required for persistent infection and transmission by Frankliniella occidentalis. J. Virol. 2014, 88, 5788–5802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Shi, M.; Lin, X.D.; Tian, J.H.; Chen, L.J.; Chen, X.; Li, C.X.; Qin, X.-C.; Li, J.; Cao, J.-P.; Eden, J.-S.; et al. Redefining the invertebrate RNA virosphere. Nature 2016, 540, 539–543. [Google Scholar] [CrossRef] [PubMed]
  37. Abudurexiti, A.; Adkins, S.; Alioto, D.; Alkhovsky, S.V.; Avšič-Županc, T.; Ballinger, M.J. Taxonomy of the order Bunyavirales: Update 2019. Arch. Virol. 2019, 164, 1949–1965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. 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] [PubMed]
  39. 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] [Green Version]
  40. 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]
  41. 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] [PubMed]
  42. 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] [PubMed]
  43. 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] [PubMed]
  44. 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] [PubMed]
  45. Nerva, L.; Forgia, M.; Ciuffo, M.; Chitarra, W.; Chiapello, M.; Vallino, M.; Varese, G.C.; Turina, M. The mycovirome of a fungal collection from the sea cucumber Holothuria polii. Virus Res. 2019, 273, 197737. [Google Scholar] [CrossRef] [PubMed]
  46. Sasai, S.; Tamura, K.; Tojo, M.; Herrero, M.L.; Hoshino, T.; Ohki, S.T.; Mochizuki, T. A novel non-segmented double-stranded RNA virus from an Arctic isolate of Pythium polare. Virology 2018, 522, 234–243. [Google Scholar] [CrossRef] [PubMed]
  47. Botella, L.; Janousek, J.; Maia, C.; Horta Jung, M.; Raco, M.; Jung, T. 2020. Marine Oomycetes of the genus Halophytophthora harbor viruses related to Bunyaviruses. Front. Microbiol. 2020, 11, 1467. [Google Scholar] [CrossRef] [PubMed]
  48. Chiapello, M.; Rodríguez-Romero, J.; Ayllón, M.A.; Turina, M. Analysis of the virome associated to grapevine downy mildew lesions reveals new mycovirus lineages. Virus Evol. 2020, 6, veaa058. [Google Scholar] [CrossRef] [PubMed]
  49. 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]
  50. Anagnostakis, S.L.; Day, P.R. Hypovirulence conversion in Endothia parasitica. Phytopathology 1979, 69, 1226–1229. [Google Scholar] [CrossRef]
  51. Thapa, V.; Turner, G.G.; Hafenstein, S.; Overton, B.E.; Vanderwolf, K.J.; Roossinck, M.J. Using a novel partitivirus in Pseudogymnoascus destructans to understand the epidemiology of white-nose syndrome. PLoS Pathog. 2016, 12, e1006076. [Google Scholar] [CrossRef] [PubMed]
  52. Michel, B.E. Evaluation of the water potentials of solutions of polyethylene glycol 8000 both in the absence and presence of other solutes. Plant Physiol. 1983, 72, 66–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Harris, A.R.; Webber, J.F. Sporulation potential, symptom expression and detection of Phytophthora ramorum on larch needles and other foliar hosts. Plant Pathol. 2016, 65, 1441–1451. [Google Scholar] [CrossRef] [Green Version]
  54. Jung, T.; Stukely, M.J.C.; Hardy, G.S.J.; White, D.; Paap, T.; Dunstan, W.A.; Burgess, T.I. Multiple new Phytophthora species from ITS Clade 6 associated with natural ecosystems in Australia: Evolutionary and ecological implications. Persoonia 2011, 26, 13–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Jung, T.; Horta Jung, M.; Scanu, B.; Seress, D.; Kovács, G.; Maia, C.; Pérez-Sierra, A.; Chang, T.-T.; Chandelier, A.; Heungens, K.; et al. Six new Phytophthora species from ITS Clade 7a including two sexually functional heterothallic hybrid species detected in natural ecosystems in Taiwan. Persoonia 2017, 38, 100–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Kallio, M.A.; Tuimala, J.T.; Hupponen, T.; Klemelä, P.; Gentile, M.; Scheinin, I.; Koski, M.; Käki, J.; Korpelainen, E.I. Chipster: User-friendly analysis software for microarray and other high-throughput data. BMC Genom. 2011, 12, 507. [Google Scholar] [CrossRef] [Green Version]
  57. Dufková, H.; Berka, M.; Greplová, M.; Shejbalová, Š.; Hampejsová, R.; Luklová, M.; Domkářová, J.; Novák, J.; Kopačka, V.; Brzobohatý, B.; et al. The Omics Hunt for Novel Molecular Markers of Resistance to Phytophthora infestans. Plants 2022, 11, 61. [Google Scholar] [CrossRef] [PubMed]
  58. Yang, M.; Duan, S.; Mei, X.; Huang, X.; Chen, W.; Liu, Y.; Guo, C.; Yang, T.; Wei, W.; Liu, X.; et al. The Phytophthora cactorum genome provides insights into the adaptation to host defense compounds and fungicides. Sci. Rep. 2018, 8, 6534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Dorfer, V.; Pichler, P.; Stranz, T.; Stadlmann, J.; Taus, T.; Winkler, S.; Mechtler, K. MS Amanda, a universal identification algorithm optimized for high accuracy tandem mass spectra. J. Proteome Res. 2014, 13, 3679–3684. [Google Scholar] [CrossRef] [PubMed]
  60. Thelen, M.P.; Wirth, B.; Kye, M.J. Mitochondrial defects in the respiratory complex I contribute to impaired translational initiation via ROS and energy homeostasis in SMA motor neurons. Acta Neuropathol. Commun. 2020, 8, 223. [Google Scholar] [CrossRef] [PubMed]
  61. Rytkönen, A.; Lilja, A.; Petäistö, R.-L.; Hantula, J. Irrigation water and stem lesions on Betula pendula in a forest nursery. Scan. J. For. Res. 2008, 23, 404–411. [Google Scholar] [CrossRef]
  62. Rytkönen, A.; Lilja, A.; Vercauteren, A.; Sirkiä, S.; Parikka, P.; Soukainen, M.; Hantula, J. Identity and potential pathogenity of Phytophthora species found on symptomatic rhododendron plants in a Finnish nursery. Can. J. Plant Pathol. 2012, 34, 255–267. [Google Scholar] [CrossRef]
  63. Muszewska, A.; Steczkiewicz, K.; Stepniewska-Dziubinska, M.; Ginalski, K. Cut-and-paste transposons in fungi with diverse lifestyles. Genome Biol. Evol. 2017, 9, 3463–3477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. McGowan, J.; Fitzpatrick, D.A. Genomic, network and phylogenetic analysis of the Oomycete effector arsenal. mSphere 2017, 2, e00408-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Gamir, J.; Darwiche, R.; van’t Hof, P.; Choudhary, V.; Stumpe, M.; Schneiter, R.; Mauch, F. The sterol-binding activity of PATHOGENESIS-RELATED PROTEIN 1 reveals the mode of action of an antimicrobial protein. Plant J. 2016, 89, 502–509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhu, H.; Situ, J.; Guan, T.; Dou, Z.; Kong, G.; Jiang, Z.; Xi, P. A C2H2 zinc finger protein PlCZF1 is necessary for oospore development and virulence in Peronophythora litchii. Int. J. Mol. Sci. 2022, 23, 2733. [Google Scholar] [CrossRef]
  67. Feng, B.Z.; Zhu, X.P.; Fu, L.; Lv, R.F.; Storey, D.; Tooley, P.; Zhang, X.G. Characterization of necrosis-inducing NLP proteins in Phytophthora capsici. BMC Plant Biol. 2014, 8, 126. [Google Scholar] [CrossRef] [Green Version]
  68. Xu, C.; Min, J. Structure and function of WD40 domain proteins. Protein Cell 2011, 2, 202–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Li, J.; Mahajan, A.; Tsai, M.-D. Ankyrin repeat: A unique motif mediating protein−protein interactions. Biochemistry 2006, 45, 15168–15178. [Google Scholar] [CrossRef] [Green Version]
  70. Zhang, B.; Zhang, Z.; Yong, S.; Yu, S.; Feng, H.; Yin, M.; Ye, W.; Wang, Y.; Qiu, M. An oomycete-specific leucine-rich repeat-containing protein is involved in zoospore flagellum development in Phytophthora sojae. Phytopathology 2022. [CrossRef] [PubMed]
  71. Kamoun, S. A catalogue of the effector secretome of plant pathogenic oomycetes. Annu. Rev. Phytopathol. 2006, 44, 41–60. [Google Scholar] [CrossRef] [PubMed]
  72. Qiao, Y.; Liu, L.; Xiong, Q.; Flores, C.; Wong, J.; Shi, J.; Wang, X.; Liu, X.; Xiang, Q.; Jiang, S.; et al. Oomycete pathogens encode RNA silencing suppressors. Nat. Gen. 2013, 45, 330–333. [Google Scholar] [CrossRef] [Green Version]
  73. Huang, J.; Gu, L.; Zhang, Y.; Yan, T.; Kong, G.; Kong, L.; Guo, B.; Qiu, M.; Wang, Y.; Jing, M.; et al. An oomycete plant pathogen reprograms host pre-mRNA splicing to subvert immunity. Nat. Commun. 2017, 8, 2051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kharel, A.; Islam, M.T.; Rookes, J.; Cahill, D. How to unravel the key functions of cryptic Oomycete elicitin proteins and their role in plant disease. Plants 2021, 10, 1201. [Google Scholar] [CrossRef] [PubMed]
  75. Derevnina, L.; Dagdas, Y.F.; De la Concepcion, J.C.; Bialas, A.; Kellner, R.; Petre, B.; Domazakis, E.; Du, J.; Wu, C.-H.; Lin, X.; et al. Nine things to know about elicitins. New Phytol. 2016, 212, 888–895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Li, W.; Zhao, D.; Dong, J.; Kong, X.; Zhang, Q.; Li, T.; Meng, Y.; Shan, W. AtRTP5 negatively regulates plant resistance to Phytophthora pathogens by modulating the biosynthesis of endogenous jasmonic acid and salicylic acid. Mol. Plant Pathol. 2020, 21, 95–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Darino, M.; Chia, K.-S.; Marques, J.; Aleksza, D.; Soto-Jiménez, L.M.; Saado, I.; Uhse, S.; Borg, M.; Betz, R.; Bindics, J.; et al. Ustilago maydis effector Jsi1 interacts with Topless corepressor, hijacking plant jasmonate/ethylene signaling. New Phytol. 2021, 229, 3393–3407. [Google Scholar] [CrossRef] [PubMed]
  78. Gan, P.H.P.; Shan, W.; Blackman, L.M.; Hardham, A. Characterization of cyclophilin-encoding genes in Phytophthora. Mol. Genet. Genom. 2009, 281, 565–578. [Google Scholar] [CrossRef]
  79. Qutob, D.; Tedman-Jones, J.; Gijzen, M. Effector-triggered immunity by the plant pathogen Phytophthora. Trends Microbiol. 2006, 14, 470–473. [Google Scholar] [CrossRef] [PubMed]
  80. Jiang, R.H.; Weide, R.; van de Vondervoort, P.J.; Govers, F. Amplification generates modular diversity at an avirulence locus in the pathogen Phytophthora. Genome Res. 2016, 16, 827–840. [Google Scholar] [CrossRef] [Green Version]
  81. Dou, D.; Kale, S.D.; Liu, T.; Tang, Q.; Wang, X.; Arredondo, F.D.; Basnayake, S.; Whisson, S.; Drenth, A.; Maclean, D.; et al. Different domains of Phytophthora sojae effector Avr4/6 are recognized by soybean resistance genes Rps 4 and Rps 6. Mol. Plant-Microbe. Interact. 2010, 23, 425–435. [Google Scholar] [CrossRef] [PubMed]
  82. Kumar, D.; Bansal, G.; Narang, A.; Basak, T.; Abbas, T.; Dash, D. Integrating transcriptome and proteome profiling: Strategies and applications. Proteomics 2016, 16, 2533–2544. [Google Scholar] [CrossRef] [PubMed]
  83. Mergner, J.; Frejno, M.; List, M.; Papacek, M.; Chen, X.; Chaudhary, A.; Samaras, P.; Richter, S.; Shikata, H.; Lang, D.; et al. Mass-spectrometry-based draft of the Arabidopsis proteome. Nature 2020, 579, 409–414. [Google Scholar] [CrossRef] [PubMed]
  84. Pánek, M.; Fér, T.; Mráček, J.; Tomšovský, M. Evolutionary relationships within the Phytophthora cactorum species complex in Europe. Fungal Biol. 2016, 120, 836–851. [Google Scholar] [CrossRef] [PubMed]
  85. Cai, G.; Hillman, B.I. Phytophthora viruses. Adv. Virus Res. 2013, 86, 327–350. [Google Scholar] [CrossRef] [PubMed]
  86. Van Poucke, K.; Haegeman, A.; Goedefroit, T.; Focquet, F.; Leus, L.; Horta Jung, M.; Nave, C.; Redondo, M.A.; Husson, C.; Kostov, K.; et al. Unravelling hybridization in Phytophthora using phylogenomics and genome size estimation. IMA Fungus 2021, 12, 16. [Google Scholar] [CrossRef]
Figure 1. Morphological features of zoospores and sporangia, and sporulation pattern of P. cactorum isolate PhF66– on clarified VA. (a) A zoospore observed after stimulated zoospore release; (b) direct proliferation of PhF66– sporangium, and (c) release of cytoplasm from the sporangia. Scale bar = 10 µm.
Figure 1. Morphological features of zoospores and sporangia, and sporulation pattern of P. cactorum isolate PhF66– on clarified VA. (a) A zoospore observed after stimulated zoospore release; (b) direct proliferation of PhF66– sporangium, and (c) release of cytoplasm from the sporangia. Scale bar = 10 µm.
Viruses 14 02596 g001
Figure 2. Fragaria × ananassa (Malling Centenary) 20 days post inoculation with Phytophthora cactorum isolates PhF17/19 (a) and PhF66 (b). (c) Betula pendula seedling six days post inoculation with P. cactorum isolate PhF66.
Figure 2. Fragaria × ananassa (Malling Centenary) 20 days post inoculation with Phytophthora cactorum isolates PhF17/19 (a) and PhF66 (b). (c) Betula pendula seedling six days post inoculation with P. cactorum isolate PhF66.
Viruses 14 02596 g002
Figure 3. Mean lesion lengths from Phytophthora cactorum pathogenicity test on silver birch at 13 dpi. Error bars: 95% CI.
Figure 3. Mean lesion lengths from Phytophthora cactorum pathogenicity test on silver birch at 13 dpi. Error bars: 95% CI.
Viruses 14 02596 g003
Table 1. Phytophthora cactorum isolates used in this study. The viruses infecting these isolates are also given. PcAEV1–3 = Phytophthora cactorum alphaendornavirus 1–3; PcBV1–3 = Phytophthora cactorum bunya-like viruses 1–3; PcRV1 = Phytophthora cactorum RNA virus 1 (putative Totiviridae; [14]); PcUV1 = Phytophthora cactorum usti-like virus (PcUV1). The isolates were used in the following analyses of this study; I = virus curing, II = virus transmission, III = virus effects (growth, sporulation, pathogenicity, transcriptome, proteome), IV = pathogenicity.
Table 1. Phytophthora cactorum isolates used in this study. The viruses infecting these isolates are also given. PcAEV1–3 = Phytophthora cactorum alphaendornavirus 1–3; PcBV1–3 = Phytophthora cactorum bunya-like viruses 1–3; PcRV1 = Phytophthora cactorum RNA virus 1 (putative Totiviridae; [14]); PcUV1 = Phytophthora cactorum usti-like virus (PcUV1). The isolates were used in the following analyses of this study; I = virus curing, II = virus transmission, III = virus effects (growth, sporulation, pathogenicity, transcriptome, proteome), IV = pathogenicity.
Isolate NameVirusesIsolation YearStrawberry CultivarCollection LocationUsed in Analyses
PhF9PcAEV1, PcAEV2, PcRV11990JonsokPohjaI
PhF38PcAEV1, PcAEV2, PcRV11993unknownIlomantsiII
PhF66PcBV1, PcBV22006PolkaNurmijärviI, II, III, IV
PhF79PcBV1, PcBV2, PcBV3, PcRV1, PcUV11991JonsokPohjaI, II
PhF101PcAEV1, PcAEV3, PcRV11993JonsokPohjaII
PhF17/19PcRV12019PolkaHuhdasjärviIV
Ph415unknown2004Betula pendulaSuonenjokiIV
Table 2. Morphological characteristics measured from the virus-free P. cactorum isolate PhF66– and PhF66 containing two bunyaviruses.
Table 2. Morphological characteristics measured from the virus-free P. cactorum isolate PhF66– and PhF66 containing two bunyaviruses.
SporangiaOogoniaAntheridia
Length/µmBreadth/µmSporangia FormAbundance (№ per Agar Plug)Length
/µm
Breadth
/µm
Oogonia
Form
Oospore Diam./µmOospore Wall/µmLength
/µm
Breadth
/µm
Antheridia
Form
Abortion Rate
PhF66Mean
Min
Max
36.37
26.35
45.11
29.77
22.14
37.05
broad-ovoid and ovoid, also spherical897
453
599
32.49
27.19
39.09
30.60
24.16
35.24
smooth, globous to sub-globous25.87
21.57
30.79
1.27
0.44
1.83
14.52
9.55
26.41
10.50
7.01
23.52
Paragynus, clavate29.27%
20%
38%
PhF66–Mean
Min
Max
45.10
33.87
57.13
35.58
29.72
42.75
ovoid to slight limoniform519
724
1148
33.66
28.24
39.12
29.81
24.57
33.64
smooth, globous to sub-globous25.81
21.57
32.81
1.33
0.84
1.84
14.04
9.53
22.88
9.98
5.86
12.51
Paragynus, clavate91.11%
85%
97%
N 5050502076/3588 (PhF66/PhF66–)5050505050505050100
p-value sig. level ≤0.00625 4.207 × 10−131.135 × 10−12 0.0080730.032950.06979 0.87060.22270.44110.2362 0.000586
Table 3. Differentially expressed genes in the virus free P. cactorum isolate PhF66– and PhF66 containing two bunyaviruses with cutoff values >2 revealed by RNA-seq. * = corresponding proteins detected differentially abundant alsoin the shotgun proteomic analysis. ** = corresponding proteins detected, but not showing differences in abundance. Annotated elicitins are highlighted in bold.
Table 3. Differentially expressed genes in the virus free P. cactorum isolate PhF66– and PhF66 containing two bunyaviruses with cutoff values >2 revealed by RNA-seq. * = corresponding proteins detected differentially abundant alsoin the shotgun proteomic analysis. ** = corresponding proteins detected, but not showing differences in abundance. Annotated elicitins are highlighted in bold.
Read Counts per Million
PhF66– (No Viruses) PhF66 (PcBV1&PcBV2)
AccessionDescriptionLengthR1R2R3R7R8R4R5R6R9R10log2Fold ChangeFDRPadj
KAF1787014.1hypothetical protein GQ600_183325810.000.000.000.000.0012.9117.7821.2721.1915.596.560.000002<0.001
KAF1778147.1Riboflavin synthase-like beta-barrel14441.500.921.544.681.6622.0854.908.288.2810.262.700.000053<0.001
KAF1774323.1Elicitin53452.6347.3147.617.908.5745.4377.86283.71293.98292.022.680<0.001
KAF1774322.1Elicitin5160.110.160.000.080.000.070.252.922.772.322.490.000513<0.001
KAF1789104.1hypothetical protein GQ600_99252700.750.840.350.450.584.825.014.022.773.872.430.00005<0.001
KAF1774139.1hypothetical protein GQ600_100434700.150.080.040.120.120.912.960.691.261.222.390.000147<0.001
KAF1781019.1 *Elicitin3571.940.721.410.040.000.981.346.9928.6313.272.350.000294<0.001
KAF1780261.1SMP-30/Gluconolactonase/LRE-like region11142.132.321.192.651.1215.5914.538.437.827.162.190.000161<0.001
KAF1780813.1hypothetical protein GQ600_5602641.311.640.840.900.427.402.347.395.936.802.140.000655<0.001
KAF1785474.1Nucleotide-diphospho-sugar transferase8790.820.320.662.280.876.9619.452.432.522.772.070.000655<0.001
KAF1788825.1Ankyrin repeat-containing domain117434.0938.8336.0133.3139.710.000.000.250.290.24−3.700<0.001
KAF1791464.1WD40/YVTN repeat-like-containing domain57995.274.965.474.565.410.360.040.300.500.45−3.240.000041<0.001
KAF1788826.1hypothetical protein GQ600_30507622.731.961.940.491.700.000.000.000.000.00−3.210.000421<0.001
KAF1775491.1WD40-repeat-containing domain62382.772.202.602.521.580.110.380.100.040.08−2.720.000065<0.001
KAF1791150.1hypothetical protein GQ600_2466831629.7911.8012.5813.9713.023.081.540.690.760.85−2.690.000062<0.001
KAF1789612.1putative domain, di-copper centre171215.5912.7613.945.334.950.980.382.281.891.83−2.540.000029<0.001
KAF1790519.1putative domain, di-copper centre4442.802.643.351.180.750.220.000.300.130.28−2.450.000347<0.001
KAF1789787.1hypothetical protein GQ600_243243114.453.723.792.001.040.250.210.500.630.16−2.340.000161<0.001
KAF1792605.1hypothetical protein GQ600_23700640311.5113.0811.2129.0723.204.503.422.332.401.91−2.320.000008<0.001
KAF1795150.1Tetratricopeptide repeat722216.1515.1616.0230.8624.704.684.973.572.482.61−2.310.000001<0.001
KAF1782989.1WD repeat-containing protein 9075155.054.925.475.134.321.201.130.450.670.28−2.290.000164<0.001
KAF1782797.1 **Thrombospondin type-1 (TSP1) repeat718467.6256.0758.4214.6611.351.992.4612.7411.949.53−2.290.000001<0.001
KAF1783174.1hypothetical protein GQ600_255205580.780.840.880.410.710.000.000.000.040.04−2.280.000328<0.001
KAF1775375.1 **Carbonic anhydrase, alpha-class, conserved site66124.4520.7623.171.551.750.510.751.742.771.83−2.240.000232<0.001
KAF1793614.1hypothetical protein GQ600_1916839719.389.609.5810.4213.061.811.342.581.851.55−2.230.00003<0.001
KAF1783504.1WD40-repeat-containing domain30391.162.121.721.592.490.330.380.200.130.00−2.210.000635<0.001
KAF1783853.1Katanin p60 subunit A1209111.8515.6414.2112.5414.973.443.421.981.642.48−2.190.000019<0.001
KAF1793060.1EF-Hand 1, calcium-binding site8041.461.921.851.260.620.040.130.100.170.33−2.180.000247<0.001
KAF1772692.1EF-hand domain pair27391.311.201.941.340.830.150.130.000.000.00−2.120.000757<0.001
KAF1789078.1WD40-repeat-containing domain29612.803.763.223.914.120.801.040.250.380.69−2.120.000098<0.001
KAF1794169.1Organic solute carrier protein 153035.536.966.3511.8110.353.373.460.790.420.28−2.110.000655<0.001
KAF1782147.1P-loop containing nucleoside triphosphate hydrolase136326.655.244.857.948.481.342.001.690.630.69−2.100.000126<0.001
KAF1788166.1hypothetical protein GQ600_166216740.520.600.260.900.500.000.000.000.000.00−2.040.002536<0.001
Table 4. Differentially abundant proteins (p < 0.05) in the virus free PhF66– (samples S2–4) and PhF66 (samples S5–7) containing two bunyaviruses by RNA-seq. * = proteins also found differentially expressed in RNA-seq (p < 0.05). “High” and “Peak Found” represent proteins identified by matching peptide MS/MS spectra and proteins that were identified by matching combinations of expected peptide mass and retention time, respectively. Annotated elicitins are highlighted in bold.
Table 4. Differentially abundant proteins (p < 0.05) in the virus free PhF66– (samples S2–4) and PhF66 (samples S5–7) containing two bunyaviruses by RNA-seq. * = proteins also found differentially expressed in RNA-seq (p < 0.05). “High” and “Peak Found” represent proteins identified by matching peptide MS/MS spectra and proteins that were identified by matching combinations of expected peptide mass and retention time, respectively. Annotated elicitins are highlighted in bold.
PhF66– (No Viruses)PhF66 (PcBV1&PcBV2)
AccessionDescriptionUnique PeptidesBlankS2S3S4S5S6S7log2Fold ChangePadj
KAF1782669.1S-adenosyl-L-methionine-dependent methyltransferase2Not FoundNot FoundNot FoundNot FoundHighHighPeak Found6.64<0.001
KAF1784643.1P-loop containing nucleoside triphosphate hydrolase2Peak FoundNot FoundNot FoundNot FoundNot FoundHighPeak Found6.64<0.001
KAF1791085.1hypothetical protein GQ600_70935Not FoundNot FoundPeak FoundPeak FoundHighHighHigh4.69<0.001
KAF1785983.1FAD/NAD(P)-binding domain2Not FoundHighNot FoundHighNot FoundNot FoundNot Found−6.64<0.001
KAF1784304.1Ubiquitin-activating enzyme E1, Cys active site2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1777841.1Zinc finger, RING/FYVE/PHD-type2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1780096.150S ribosomal protein L30e-like2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1778206.1Nucleotide-binding alpha-beta plait domain2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1772484.1Exosome complex RNA-binding protein 1/RRP40/RRP42Peak FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1791972.1 *Zinc finger, NHR/GATA-type2Not FoundHighHighHighNot FoundNot FoundNot Found−6.64<0.001
KAF1794225.1thiamine diphosphate-binding fold2Not FoundNot FoundNot FoundNot FoundHighHighNot Found6.64<0.001
KAF1775607.1 *DNA polymerase family X lyase domain2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1772539.1Alpha/Beta hydrolase fold2Not FoundNot FoundNot FoundNot FoundNot FoundHighNot Found6.64<0.001
KAF1774000.1hypothetical protein GQ600_36472Not FoundNot FoundNot FoundNot FoundHighHighNot Found6.64<0.001
KAF1793066.1Aldolase-type TIM barrel2Not FoundPeak FoundNot FoundHighNot FoundNot FoundNot Found−6.64<0.001
KAF1795828.1Late embryogenesis abundant protein LEA4Not FoundPeak FoundHighNot FoundHighHighHigh3.08<0.001
KAF1790797.1hypothetical protein GQ600_245924Not FoundPeak FoundHighPeak FoundHighHighHigh3.06<0.001
KAF1781808.1Leucine-rich repeat domain, L domain-like3Not FoundNot FoundNot FoundPeak FoundNot FoundHighNot Found2.51<0.001
KAF1790245.1 *Fatty acid desaturase domain2Not FoundNot FoundNot FoundHighNot FoundPeak FoundPeak Found−2.34<0.001
KAF1772416.1S-adenosyl-L-methionine-dependent methyltransferase2Not FoundHighHighHighPeak FoundPeak FoundPeak Found−2.15<0.001
KAF1781019.1 *Elicitin3Not FoundPeak FoundNot FoundNot FoundHighHighHigh2.30<0.001
KAF1780998.1Molybdenum cofactor biosynthesis, conserved site2Not FoundPeak FoundNot FoundPeak FoundNot FoundHighNot Found2.22<0.001
KAF1779922.1hypothetical protein GQ600_69172Not FoundNot FoundPeak FoundNot FoundHighHighPeak Found2.08<0.001
KAF1779895.1Multicopper oxidase, type 32Not FoundNot FoundPeak FoundNot FoundHighHighHigh2.03<0.001
KAF1794156.1NAD(P)-binding domain2Not FoundPeak FoundNot FoundPeak FoundNot FoundHighNot Found2.00<0.001
KAF1783208.1FAD/NAD(P)-binding domain11Peak FoundHighHighHighPeak FoundNot FoundPeak Found−1.83<0.001
KAF1789511.1TmcA/NAT10/Kre332Not FoundNot FoundHighPeak FoundHighHighHigh1.91<0.001
KAF1779266.1 *Major facilitator superfamily domain2Not FoundPeak FoundPeak FoundHighNot FoundHighPeak Found−1.73<0.001
KAF1794053.1NAD(P)-binding domain2Not FoundPeak FoundPeak FoundNot FoundHighHighPeak Found1.88<0.001
KAF1791322.1Short-chain dehydrogenase/reductase, conserved site10Not FoundHighHighHighHighHighHigh−1.720.001
KAF1787880.1Multicopper oxidase, type 33Not FoundNot FoundPeak FoundNot FoundHighHighHigh1.840.002
KAF1780390.1Ubiquitin conjugation factor E4, core2Peak FoundPeak FoundNot FoundPeak FoundPeak FoundHighPeak Found1.830.002
KAF1779223.1 *Glycoside hydrolase superfamily6Peak FoundHighHighHighHighHighHigh1.830.002
KAF1777766.1Elicitin6Not FoundPeak FoundHighPeak FoundHighHighHigh1.800.002
KAF1773149.1Thioredoxin-like fold5Not FoundHighHighHighHighHighHigh1.730.003
KAF1780973.1Leucine-rich repeat domain, L domain-like2Not FoundPeak FoundNot FoundPeak FoundNot FoundHighPeak Found1.710.004
KAF1793851.1Ubiquitin-conjugating enzyme, active site20Peak FoundHighHighHighHighHighHigh1.660.006
KAF1791020.1NAD(P)-binding domain11Peak FoundHighHighHighHighHighHigh−1.450.009
KAF1781134.1P-loop containing nucleoside triphosphate hydrolase2Not FoundPeak FoundPeak FoundHighPeak FoundNot FoundNot Found−1.420.011
KAF1791905.1FAD/NAD(P)-binding domain2Not FoundHighPeak FoundHighPeak FoundPeak FoundPeak Found−1.370.016
KAF1788119.1S-adenosyl-L-methionine-dependent methyltransferase2Not FoundHighPeak FoundHighPeak FoundPeak FoundPeak Found−1.360.018
KAF1791684.1ABC-transporter extension domain4Not FoundHighHighHighPeak FoundPeak FoundPeak Found−1.330.021
KAF1794255.1Peptidase C1A4Not FoundPeak FoundHighPeak FoundHighHighHigh1.480.022
KAF1773151.1Acyl-CoA dehydrogenase/oxidase, N-terminal13Peak FoundHighHighPeak FoundHighHighHigh1.480.023
KAF1786762.1HAD-like domain2Not FoundNot FoundPeak FoundNot FoundHighHighHigh1.460.026
KAF1794937.1Phosphoenolpyruvate carboxykinase (ATP), conserved site3Peak FoundHighHighHighPeak FoundHighHigh−1.270.032
KAF1773405.1Extradiol ring-cleavage dioxygenase, class III enzyme, subunit B3Not FoundHighNot FoundHighNot FoundNot FoundPeak Found−1.240.037
KAF1773588.1Glyoxalase/Bleomycin resistance protein/Dihydroxybiphenyl dioxygenase2Not FoundPeak FoundNot FoundNot FoundPeak FoundHighPeak Found1.400.038
KAF1782519.1ClpP/crotonase-like domain14Peak FoundHighHighHighHighHighHigh1.380.043
KAF1775431.1Cell division protein FtsZ, C-terminal3Not FoundPeak FoundPeak FoundPeak FoundPeak FoundHighPeak Found1.370.046
KAF1778058.1Protein argonaute, N-terminal10Peak FoundPeak FoundPeak FoundPeak FoundHighHighHigh1.370.046
KAF1772871.1Glycine cleavage T-protein/YgfZ, C-terminal domain2Not FoundHighHighHighPeak FoundPeak FoundHigh−1.190.05
Table 5. Increments of healthy leaves in garden strawberry plants after inoculation with Phytophthora cactorum strains in the pathogenicity test II. dpi = days post inoculation.
Table 5. Increments of healthy leaves in garden strawberry plants after inoculation with Phytophthora cactorum strains in the pathogenicity test II. dpi = days post inoculation.
TreatmentReplicate    0 dpi           20 dpi       Increase in Amount of Healthy LeavesAverage
Healthy LeavesHealthy LeavesWiltingDead
Control158003
236003
348004
4360033.25
PhF66147003
259004
335002
4460022.75
PhF66–157002
258003
337004
4360033
PhF17/1915712−1
24722−1
35752−5
44651−4
54704−1
66707−6−3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Poimala, A.; Raco, M.; Haikonen, T.; Černý, M.; Parikka, P.; Hantula, J.; Vainio, E.J. Bunyaviruses Affect Growth, Sporulation, and Elicitin Production in Phytophthora cactorum. Viruses 2022, 14, 2596. https://doi.org/10.3390/v14122596

AMA Style

Poimala A, Raco M, Haikonen T, Černý M, Parikka P, Hantula J, Vainio EJ. Bunyaviruses Affect Growth, Sporulation, and Elicitin Production in Phytophthora cactorum. Viruses. 2022; 14(12):2596. https://doi.org/10.3390/v14122596

Chicago/Turabian Style

Poimala, Anna, Milica Raco, Tuuli Haikonen, Martin Černý, Päivi Parikka, Jarkko Hantula, and Eeva J. Vainio. 2022. "Bunyaviruses Affect Growth, Sporulation, and Elicitin Production in Phytophthora cactorum" Viruses 14, no. 12: 2596. https://doi.org/10.3390/v14122596

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