The Role of Fungal Secondary Metabolites and sRNAs: Comparison
Please note this is a comparison between Version 3 by Johannes Mapuranga and Version 2 by Jessie Wu.

Fungal plant pathogens use proteinaceous effectors as well as newly identified secondary metabolites (SMs) and small non-coding RNA (sRNA) effectors to manipulate the host plant’s defense system via diverse plant cell compartments, distinct organelles, and many host genes. However, most molecular studies of plant–fungal interactions have focused on secreted effector proteins without exploring the possibly equivalent functions performed by fungal (SMs) and sRNAs, which are collectively known as “non-proteinaceous effectors”. Fungal SMs have been shown to be generated throughout the plant colonization process, particularly in the early biotrophic stages of infection. The fungal repertoire of non-proteinaceous effectors has been broadened by the discovery of fungal sRNAs that specifically target plant genes involved in resistance and defense responses. Many RNAs, particularly sRNAs involved in gene silencing, have been shown to transmit bidirectionally between fungal pathogens and their hosts.

  • SMs
  • sRNAs
  • interactions

1. Introduction

Plants have developed a broad spectrum of responses to counter pathogen invasion. Likewise, plant pathogens orchestrate a highly calibrated array of pathogenicity strategies in their quest to cause diseases [1]. The recent increased availability of fungal and plant genomes in the public domain has facilitated considerable progress in molecular plant–fungal interaction studies. Using genetic techniques, pathogenicity or virulence factors have been established, and the study of these factors has increased our understanding of the interactions between pathogens and their hosts. During interaction with their hosts, fungal plant pathogens secrete many proteins known as effectors which manipulate the physiology of the host or suppress the host’s immunity to promote infection [2][3]. Most studies on effectors have focused almost exclusively on secreted proteins, without exploring the possibly equivalent functions performed by fungal secondary metabolites (SMs) (chemical effectors) and sRNAs (sRNA effectors) which are collectively referred to as non-proteinaceous effectors [2][4]. Accumulating evidence has indicated that, pathogens use sRNAs (such as siRNAs and microRNAs) and SMs to manipulate host cell functions [5][6][7][8][9]. Fungal SMs and sRNAs have been shown to manipulate host defense-related genes in the same was as proteinaceous effectors [7][10][11]. In general, SM and sRNA effectors are increasingly becoming important targets for studying the pathogenesis mechanisms of fungal pathogens [11]. Furthermore, these recently discovered SM and sRNA effector entities have been shown in a number of studies to be essential in manipulating host immunity and defense-related genes [2]. It is thus important to adopt new experimental methods to elucidate the in-planta biology of SM and sRNA effectors. 

2. Fungal SMs Enhance Pathogenicity during Plant-Fungal Pathogen Interactions 

Fungal SMs are not required for the growth and development of the fungus, but they have the potential to improve the pathogen’s fitness under certain conditions. Fungal SMs are often divided into polyketides, terpenes, non-ribosomal peptides and alkaloids on the basis of the primary enzymes and precursors that are involved in their biosynthesis [12][13][14]. They play a role prior to disease by shaping the plant microbial community, allowing producers to be fully adapted. The existence of fungal SMs, which have no discernible effect on the viability of the producer, raises issues about their potential influence on the environment [15]. SMs production by fungal pathogens and the presence of a host protein that is specifically susceptible to the corresponding toxin determines the ability of the pathogen to infect the host plant. Because host-specific toxin targets are encoded by plant genes, such genes can be referred to as dominant susceptibility genes [16]. It is generally known that mutualistic or pathogenic interactions between plants and fungal pathogens entail the simultaneous generation of molecular signals [10][17].

Accumulating evidence indicates that fungal SMs serve as avirulence factors, host defense suppressors, and fungal cell wall hardening factors [10][18]. Fungal SMs are most effective during the early stage of infection (biotrophic phase), enhancing the fungus’ ability to penetrate and colonize its host without killing its host [2]. Fungal SMs can be host specific or non-host specific and generate necrosis in plant tissue. However, some fungal SMs have functions linked to virulence that are not related to necrosis [1][19][20]. As long as the host plant has the relevant molecular target, such as a resistance gene product, SMs serving as host-specific effectors are thought to play an important role in pathogen virulence [1][21]. Paradoxically, SMs acting as non-host specific effectors have been widely regarded as critical components of pathogen arsenals, despite the fact that they may not be required for pathogenesis [22]. The majority of the fungal SMs have not been defined chemically, and the plants that they are intended to affect are still a mystery. The biological actions that have been reported to be caused by fungal SMs generated in-planta suggest that they have a broad range of plant cellular targets. Some fungi use high affinity iron chelator siderophores synthesized by NRPSs to scavenge environmental iron or to sequester cellular reactive iron [23][24]. These siderophores are essential for fungal growth and development, thus enhancing pathogenicity of various fungal pathogens. Cytochalasans, a diverse group of fungal PKS-NRPS hybrid metabolites, inhibit actin polymerization [25]. The production and tranproduction and transport of proteins are targets of a wide range of fungal SMs [2619][2720]. For example, the mycotoxin deoxynivalenol (DON), a member of the type B trichothecenes, produced by Fusarium spp., inhibits protein biosynthesis by binding to the ribosome, resulting in cell signaling, differentiation, reproduction and even teratogenicity disorders in eukaryotes [2821][2922][3023]. A comparative transcriptome analysis of symptomatic and symptomless wheat tissues revealed a substantial induction of TRI genes in symptomless tissues, indicating that DON plays an important role in modulating host defenses and infection establishment [31]. Metabolite profiling of F. graminearum wild-type and the tri5 deletion mutant in infected rachis nodes supports the function of DON in suppressing host defense-related metabolites [32]. DON was demonstrated to modulate programmed cell death (PCD) of host plant cells in a concentration-dependent way [33][34]. A higher concentration of DON may be produced by F. graminearum during infection to trigger hydrogen peroxide (H2O2) production by increasing the size of the hyphal colony. This results in further induction of PCD in wheat [33][34], and thus enhances its switch from biotrophy to necrotrophy [32]. Therefore, it can be unequivocally concluded that, during infection, the mycotoxin DON is produced as a sophisticated strategy of the fungal pathogen to circumvent and hijack the host plant’s defense system. 

Cochliobolus species were reported to produce host-specific toxins that enhance pathogen virulence. Victorin, a non-ribosomal peptide produced by Cochliobolus victoriae, is a virulence factor that enhances pathogenicity by inducing PCD during infection of only oat cultivars harboring susceptible genes [3524][3625][3726]. It was also reported that victorin targets the plasma membrane and triggers PCD signaling pathways. HC-toxin, a non-ribosomal peptide produced by Cochliobolus carbonum induces histone hyperacetylation through the inhibition of histone deacetylases, during the infection of only maize varieties harboring susceptible genes [3827][3928]. Transcriptional activation of host plant defense genes is altered by such histone modifications, thereby enhancing pathogen virulence [3928][4029]. Alternaria alternata have various pathotypes that produce different host specific toxins that are active only on their corresponding susceptible hosts [4130]. Some host specific toxins including destruxin which is produced by Alternaria brassicae are also essential for pathogens in susceptible host plants. A. alternata also produces AAL-toxin, an SM which enhances pathogenicity in tomato varieties harboring susceptible genes by inhibiting ceramide synthase. This will lead to free phytosphingosine and sphinganine accumulation followed by the disruption of plasma membrane [4231]. Depudecin is another SM produced by Alternaria brassicicola which also enhances pathogenicity by inhibiting histone deacetylases; its role in pathogenicity is weaker than that of HC-toxin [4332]. Tenuazonic acid produced by members of the genus Alternaria and other phytopathogenic fungi inhibits protein biosynthesis on ribosomes [44][45]. It was hypothesized that the Colletotrichum graminicola disease cycle is supported by monorden and monocillins in various ways, initially promoting biotrophic asymptomatic infection by inhibiting Hsp90 chaperons of R-proteins, and disrupting a maize hypersensitive response by enabling a switch to necrotrophy through suppression of basal plant defenses [46]. Fungal SMs such as ophiobolin and herbarumin enhance virulence by inhibiting calmodulin signaling which will disrupt plant regulatory networks [47][48]. It was also demonstrated that Colletotrichum higginsianin SM higginsianin B inhibits jasmonate-mediated plant defenses [49]. Two non-ribosomal octapeptides Fusaoctaxin A and B, which are biosynthesized by the gene cluster fg3_54, were found to be F. graminearum virulence factors [50][51]. Fusaoctaxin A alters the subcellular localization of chloroplasts in coleoptile cells and inhibits callose deposition in plasmodesmata during pathogen infection, thereby facilitating F. graminearum cell-to-cell penetration in wheat cells [50][51]. 

3. sRNAs - The Secret agents in Plant-Fungal Pathogen Interactions 

Plant immune responses are tightly regulated by an array of immunity-associated regulators such as sRNAs and some transcription factors [5233]. Based on their biogenesis and structural features, sRNAs can be classified into three categories: short-interfering RNAs (siRNAs), dicer-independent microRNAs (miRNAs) and dicer-independent piwi interacting RNAs (piRNAs) [9][5334][5435][5536]. The fundamental sRNA pathway components and other various sRNAs function as critical gene expression regulators to fine-tune the immunity of some cereal plants such as wheat and rice against pathogen invasion [5233]. Normally, when a pathogen attacks its host, these sRNAs are either upregulated or downregulated in order to inhibit expression or to release suppression of their targets [5][5637]. Thus, plant endogenous sRNAs and sRNA pathway components play key roles in regulating and fine-tuning host immune responses to pathogens such as fungi, bacteria, and oomycetes [5738]. Accumulating evidence indicates that sRNAs produced by fungal pathogens can function as effector molecules, modulating host gene expression as a counter-defense mechanism (Table 1) [5][7][9][5839][5940][6041][6142].
Table 1. Fungal sRNA effectors and their target genes in cross-kingdom interactions.

4. Cross-Kingdom RNAi during Plant-Fungal Pathogen Interactions

During pathogen infection, the host’s sRNA performs an endogenous role by regulating gene expression in order to maintain a healthy balance between plant development and immunity [7657][7758][7859]. sRNAs can direct the transcriptional and post-transcriptional silencing of gene expression, and this phenomenon is known as RNA interference (RNAi). Post-transcriptional gene silencing is a mechanism through which plant miRNAs contribute to resistance by regulating the expression of defense-related genes [7960][8061]. The phenomena of cross-kingdom RNAi occurs when gene silencing is induced between unrelated species from different kingdoms, like a plant host and its interacting pathogen (Figure 1). It necessitates the translocation of a gene-silencing trigger from a donor into an interacting recipient. Several studies have reported interactions with other species in plant and animal systems through cross-kingdom RNAi [5738][8162][8263]. sRNAs generated by pathogens and parasites, on the other hand, may also translocate into host cells and induce host gene silencing [5839][8162][8263][8364][8465]. This implies that sRNA transfer is bidirectional; plant-derived sRNAs serve as defense weapons to disrupt fungal pathogenicity genes, while pathogen-derived sRNAs act as offensive weapons to suppress host plant defense mechanisms.
Figure 1. Cross-kingdom RNAi and vesicle trafficking during plant-fungal pathogen interactions. Fungal and plant sRNAs trigger cross-kingdom RNAi during plant-pathogen interactions. Fungal sRNAs translocate into plant cells and hijack the host plant Argonaute (AGO) protein of the RNAi machinery to suppress host plant immune response. The fungal sRNAs are upregulated upon infection (indicated by green arrow). Host cells also can deliver sRNAs into pathogen cell, either host induced gene silencing (HIGS) sRNAs or endogenous sRNAs, to target virulence genes and other essential pathogen genes. The generation of multivesicular bodies and release of exosomes at the site of pathogen invasion is part of the host penetration resistance pathway. Among other molecules, the putative exosomes contain sRNAs that can target vesicle trafficking components of the pathogen. Exosomes can also inhibit fungal growth and stall further ingress. The production of pathogen-derived sRNAs that may target and silence host genes can be inhibited by this form of host plant immunity. The fungal pathogens also secrete proteinaceous effectors through the haustorium into the host cells to suppress the host immunity genes, thereby causing disease. How fungal pathogens transport proteinaceous effectors and sRNAs into their host cells is still elusive. On the other hand, plants secrete extracellular vesicles to transport host sRNAs into pathogens to silence fungal genes involved in pathogenicity. Passage of host sRNAs through the haustorial cell wall, either active or passive, occurs and once inside the fungal haustorium the silencing molecules trigger RNAi of their mRNA targets, and may act as primers in the fungal silencing pathway, leading to the generation of systemic silencing signals. Cell structures are not drawn to scale.
Since the discovery of RNAi in Neurospora, sRNAs from numerous fungal species have been studied [8566]. The most well-known example of cross-kingdom RNAi from a plant to its interacting pathogen is HIGS, which occurs when a plant-produced RNAi signal triggers the silencing of a pathogen gene [8667][8768]. RNase III-like endonucleases known as Dicers produce sRNAs from hairpin-structured or double-stranded RNA [8869]. The mature sRNAs are loaded into AGO proteins to form the RNA-induced silencing complex (RISC) [5][8970]. The RISC is responsible for silencing genes that contain sequences complementary to sRNAs. By using the component of the host RNAi machinery known as AGO1, the transfer of B. cinerea sRNA into Arabidopsis cells silenced the host’s immune genes [5839]. Fungal sRNAs can suppress host plant immunity by interfering with the RNAi pathways of the host [7][5839][5940]. During tomato and Arabidopsis infection, the most prevalent sRNAs that function as effectors factors to enhance pathogen virulence are Bc-siR3.1, Bc-siR3.2, and Bc-siR3.5 which target the host mitogen-activated protein kinases MPK1, MPK2 and MPKKK4, peroxiredoxin (PRXIIF), and cell wall-associated kinase (WAK), respectively [5839]. These pathogen-derived sRNAs target components of host plant immunity such as oxidative burst and signal transduction pathways; hence, silencing of these targets will enhance pathogen virulence and compromise resistance to the fungal pathogens [5839]. In support of the hypothesis that sRNAs enhance B. cinerea virulence, a dcl1 dcl2 B. cinerea double mutant that could not produce Bc-sRNAs exhibited attenuated virulence, whilst a dcl1 or dcl2 single mutant still produced sRNAs to sustain pathogenicity on host plants [5839][9071]. B. cinerea also delivers Bc-siR37 into the host cells, which targets host PMR6, WRKY7 and FEI2, leading to suppression of host immunity [5940].
A group of fungal in-planta secreted sRNAs was also identified from the sequencing of sRNAs from Sclerotinia sclerotiorum during infection of Arabidopsis and Phaseolus vulgaris [9172]. The pathogen-derived sRNAs were predicted to target quantitative disease resistance-associated genes of the host and suppress host plant immunity [9172]. Mutations of two sRNA targets that encode kinase genes SERK2 and SNAK2 enhanced pathogen virulence and compromised host plant resistance, indicating that the sRNAs’ targets are involved in disease resistance [9172]. Analysis of Pst-infected leaves established that Pst is capable of suppressing the host’s defense and immunity genes as well as its endogenous genes by producing many sRNAs [6]. When a fungal infection occurs, the miRNAs involved in disease response either up- or down-regulate the expression of their target genes [92][93]. For example, there was a substantial increase in miR1138 levels in bread wheat infected with Pgt (62G29-1) [93]. miR393, miR444, miR827, and miR2005 were uprengulated in wheat (T. aestivum L.) following B. graminis infection [94]. The open reading frame or untranslated regions of certain genes are targeted by miRNAs, which have the ability to inhibit the translation of those genes [95]. Fungal sRNAs can ta can target and silence plant transcripts involved in defense, but sRNAs from plants can target and silence transcripts produced by pathogens [9673][9774]. A novel Pst miRNA (Pst-milR1) participates in cross-kingdom RNAi events in wheat by binding the pathogenesis-related 2 (PR2) gene, which may suppress the host-mediated defense mechanism in its counter defense. Silencing of the Pst-milR1 precursor using host induced gene silencing resulted in reduced Pst virulence and increased wheat resistance to the Pst isolate CRY31. Therefore, Pst-milR1 is a key pathogenicity factor in Pst, which functions as an effector to suppress host immunity [7]. Computational prediction of targets using a common set of sRNAs and Pt mil-RNAs (pt-mil-RNAs) within Pt and wheat found that the majority of the targets of Pt-derived sRNAs were repetitive elements in Pt, whilst in wheat the target genes were revealed to be involved in various biological processes including defense-related pathways [9]. The sRNAs’ targeted genes are involved in disease resistance, metabolic processes, transporter, and apoptotic inhibitor activities. This was the first study to report the discovery of new sRNAs found in Pt [9]. Expression validation studies performed on twenty individual Pt-sRNAs and two pt-mil-RNAs, namely pt-mil-RNA1 and pt-mil-RNA2, showed evidence of their possible role in pathogenicity or virulence on the host. pt-mil-RNA1 and pt-mil-RNA2 were both found to suppress wheat defense response to Pt by targeting wheat transcription factor TCP14, cytochrome b5 reductase and elongation factor 2 [9]. Fg-sRNA1 produced by F. graminearum targets and silences wheat TaCEBiP (Chitin Elicitor Binding Protein), a pattern recognition receptor gene (Figure 2) [6041]. F. oxysporum f. sp. lycopersici produces Fol-milR1, an sRNA effector that suppresses host immunity by targeting the tomato protein kinase SlyFRG4 via AGO4a, thus providing a novel pathogenicity strategy to achieve infection [6142].
Figure 2. Mode of action of F. graminearum-secreted pathogenicity factors during F. graminearum-wheat interaction. The sRNA effector Fg-sRNA1 contributes to virulence by silencing wheat defense-related TaCEBiP. Fungal toxin DON inhibits protein biosynthesis by binding to the ribosome. The fungal toxin Fusaoctaxin A changes the subcellular localization of chloroplasts in the coleoptile cells and prevents callose accumulation in plasmodesmata during pathogen infection, facilitating the cell-to-cell invasion of F. graminearum in wheat tissues.
Centromeric sRNAs associated with genome-wide hypermethylation were induced by the stem rust pathogen Pgt during la recente infection stages [98]. Althostugh Pgt-derived sRNAs maybe be used by the pathogen for silencing target host genes, endogenous functions of these sRNAs were also discovered during infection [98]. A recent study esta established that, upon infection, particular wheat 24-nt heterochromatic siRNAs (hc-siRNA) were repressed, whereas specific 25-nt rRNA and tRNA fragments were significantly upregulated. Transcripts encoding a ribosomal protein and a glycosyl hydrolase effector in the fungal pathogen were cleaved by wheat sRNAs [9975]. Long inverted repeats in the vicinity of protein coding genes in fungi gave rise to the development of miRNA-like and phased 21-nt sRNAs. sRNAs produced by fungi targeted not only native transcripts, such as transposons and kinases, but also transcripts from other kingdoms, such as a wheat nucleotide-binding domain leucine-rich repeat receptor (NLR) and several families of transcription factors involved in defense. This research provides new insights into host-microbe coevolution and opens up promising ways to improve biotechnology to control pathogens [9975]. Accumulating evidence shows that miRNAs serve crucial roles in regulating the expression of their target genes accurately and effectively during the interactions between rice and M. oryzae. Understanding the functions of rice miRNAs is crucial for managing rice blast. miR398b coordinates various pathways to increase the accumulation of H2O2 through numerous Superoxide Dismutase (SOD) family genes [10076], thereby positively regulating rice defense responses to M. oryzae [8061]. miR166k-miR166h, miR160a, and miR7695, positively regulate [8061][10177][10278], whereas miR444b.2, miR164a, miR319b, miR169, Osa-miR439, and miR396 negatively regulate rice resistance to M. oryzae [6748][6950][7051][7354][7455][8061]. Osa-miR167d is a member of a conserved miRNA family that functions in developmental and stress-induced responses by regulating the expression of genes encoding auxin responsive factors (ARFs). It was demonstrated that Osa-miR167d downregulates ARF12, a component of rice immunity, to enhance M. oryzae infection. Therefore, the Osa-miR167d-ARF12 regulatory module may be helpful in enhancing resistance to blast diseases [6546].
It was previously shown that plants may transfer miRNAs to the fungal pathogen Verticillium dahliae, to activate antifungal RNAi. It was recently shown that V. dahliae may secrete an effector to the plant nucleus, where it can interfere with the nuclear export of AGO1-miRNA complexes, thereby preventing antifungal RNAi and enhancing pathogen virulence. These findings revealed an antagonistic mechanism through which fungal pathogens can manipulate plant sRNA function to thwart antifungal RNAi immunity [103]. Zhu and colleagues proposed that VdSSR1 is translocated from V. dahliae cells to the nucleus of the host plant, probably through a noncanonical secretion pathway. VdSSR1 localized in the nucleus inhibits the nuclear export of AGO1-miRNA complex and mRNAs by sequestering ALY adaptors and inhibiting them from associating with the UAP56-TREX complex [103]. In the plant-V. dahliae system, VdSSR1 acts as a general suppressor of plant miRNAs, including trans-kingdom miR159 and miR166. Attenuated translocation of accumulated cytoplasmic miR166 and miR159 to V. dahliae cells ultimately suppresses trans-kingdom silencing of pathogen virulence genes and promotes fungal infection. VdSSR1-mediated suppression of the mRNA export of specific defense genes may potentially contribute to the increased virulence in plants [103]. Based on these studies, it can be concluded that sRNAs can perform a wide range of functions, including manipulating host machinery in the same way as classic proteinaceous effector molecules. Altogether, these studies demonstrate that pathogen-derived sRNAs can be translocated into the host and function as effectors to suppress host defense genes.

5. Applications of Cross-Kingdom RNAi Technology

Several studies have demonstrated RNAi-based fungal pathogen management with an average plant disease resistance of approximately 60% [10479]. Cross-kingdom RNAi was initially studied to generate disease resistance in barley and wheat against B. graminis, the powdery mildew fungus, using HIGS, an RNAi-based approach [10580]. The HIGS technique for controlling pathogens including fungi and viruses, among other plant pests, was developed with help from plant immune system’s RNA silencing machinery against viruses [10580][10681][10782]. To silence pathogen target genes, the HIGS employs RNAi by producing sequence-specific dsRNAs in the host plant. A hairpin-structure dsRNA construct which targets a specific gene is transformed into the host plant. dsRNAs and siRNAs produced by the transgenic plant are taken up by corresponding plant pathogens during host-pathogens interactions. These siRNAs target and degrade pathogen mRNAs, hence protecting the host plant against pathogen infection [8768][10883]. HIGS technology has been widely adopted in plant breeding programmes as an efficient approach to enhance plant defense responses to pathogens. The HIGS approach has been broadly utilized to manage wheat and barley fungal pathogens like the Puccinia species, Fusarium species, and B. graminis [10580][10984][11085][11186][11287][11388][11489][11590]. Obligate biotrophic fungal pathogens, Blumeria graminis f. sp. tritici (Bgt) and Bgh cause severe powdery mildew in wheat and barley, respectively. Bgh effector Avra10 and MLa10, a disease resistance gene in barley, were used to perform a proof-of-concept for HIGS [10580]. Avra10 is an important Bgt pathogenicity factor. However, recognition of Avra10 by barley MLa10 triggers hypersensitive response in the host plant, resulting in suppression of biotrophic pathogen invasion [10580][11691]. Avra10 silencing on barley leaves exhibited attenuated pathogen development without MLa10, but not with it [10580]. The HIGS approach was also used to screen fifty Bgh haustoria-associated effectors, and the silencing of eight of them significantly reduced pathogen virulence in barley [11085]. HIGS technique was further used in transgenic wheat plants to combat the rust pathogens Pst, and Pt [11388][11489][11590]. Zhu and colleagues generated transgenic wheat-derived dsRNAs targeting PsFUZ7, a MAP kinase, which contributes to Pst pathogenicity by regulating the morphology and development of hyphae [11388]. Strong resistance to Pst infection was exhibited by transgenic lines stably expressing dsRNA constructs by degrading PsFUZ7 transcripts, leading to the suppression of pathogen growth and development [11388].
It was also shown that, in transgenic wheat lines expressing target dsRNAs which disrupt the Pst cAMP-PKA chain pathway, silencing of a Pst protein kinase A (PKA) catalytic subunit (PsCK1) inhibits wheat stripe rust. Throughout the T3 and T4 generations, these transgenic lines maintained a high level of resistance to wheat stripe rust [11590]. Based on these results, PsCPK1 and its homologous genes are promising targets for developing transgenic plants with long-lasting resistance to stripe rust fugus Pst. HIGS induced by barley stripe mosaic virus (BSMV)-VIGS (virus-induced gene silencing) was established to be a robust, high-throughput approach for the functional analysis and validation of rust fungi candidate genes involved in pathogenicity through quantitative estimation of infection-related traits [11691]. It was discovered that wheat and barley plants that generated dsRNA or anti-sense RNA fragments, both of which were intended to alter gene expression in the fungus, silenced the fungal genes [10580]. Hairpin RNAi constructs with sequence similarity to MAP kinase1 (PtMAPK1) or Cyclophilin1 (PtCYC1) silenced the respective fungal genes and conferred resistance to leaf rust pathogen Pt [11489]. It was revealed that, the silencing signals in the Puccinia-wheat pathosystems are most likely host cell-derived siRNA molecules [11792]. Using site specific analysis, a recent study revealed putative cross-kingdom sRNAs, tRNA and rRNA fragments, and some signs of fungal phasing in the barley-Bgh interactions [11893]. This was the first research to report on phased short RNAs (phasiRNAs) in Bgh, a trait normally associated with plants that may be involved in the post-transcriptional regulation of fungal coding genes, pseudogenes, and transposable elements [11893].
Fusarium head blight (FHB) caused by the fungi F. graminearum and F. culmorum is another destructive disease of wheat and barley [11994][12095]. Koch and colleagues generated transgenic Arabidopsis and barley plants expressing dsRNA against three CYP51 paralogous genes that are essential in the ergosterol biosynthesis pathways [10984]. Inhibitors of sterol demethylation, which act on the CYP51 paralogs, are the most extensively used systemic fungicides for controlling fungal pathogens. Increased resistance to F. graminearum infection was shown after the transgenic expression of the three paralogous genes in two different plant systems [10984]. The expression of three hairpin RNAi constructs in wheat transgenic lines also silenced chitin synthase (Chs) 3b, a F. graminearum pathogenicity factor, and these lines exhibited firm and consistent resistance to Fusarium seedling blight (FSB) and FHB over T3-T5 generations [11186]. Fungal infection on wheat ears and seedlings was significantly reduced by the expression of dsRNAs with sequence homology to F. graminearum Chitin synthase Ch3b [11186]. Transient HIGS in wheat, and targeting an important housekeeping gene in Bg led to considerable decrease in virulence during early infection stages [7859]. Rice blast is one of the most devastating rice diseases caused by the pathogenic fungus M. oryzae [12196]. Zhu and colleagues studied the effects of silencing three virulence-related genes, ABC transporter MoABC1, membrane-bound adenylate cyclase MoMAC1, and mitogen-activated protein kinase MoPMK1, during the interactions between rice and M. oryzae [12297]. The resistance of rice to blast was enhanced when three BSMV silencing vectors targeting MoPMK1, MoMAC1, and MoABC1 were inoculated into the host plant at once. Furthermore, the development of disease was inhibited by the silencing of the individual pathogen genes [122]. Recently, a M. oryzae strain was subjected to an in vitro sensitivity assay with artificial siRN As (asiRNAs) targeting four candidate genes involved in fungal virulence: MoSSADH encoding succinic semialdehyde dehydrogenase, MoAP1 transcription factor, and its downregulated genes MoAAT encoding aminobugetyrate aminotransferase, and transcription factor MoSOM1 [123]. Feeding the fungal strain siRNAs specific to MoAP1 inhibited growth and pathogenicity, whereas siRNAs specific to MoSSADH, MoSOM1, and MoAAT did not. Additional in vivo HIGS testing showed that transgenic rice lines expressing an RNA hairpin targeting MoAP1 had enhanced resistance to eleven M. oryzae strains [123]. Altogether, cross-kingdom-RNAi has the potential to supplement existing pathogen management strategies, for instance, by widening the resistance spectra of host resistance genes. Furthermore, using cross-kingdom-RNAi to exploit naturally occurring RNA exchanges may open up new avenues for crop improvement through genetic engineering and classical breeding.

6. Translocation of Non-Proteinaceous Effectors across Kingdoms: Extracellular Vesicles as Mediators of Infection

In effector biology, the mechanisms through which fungal effectors, particularly sRNA and SM effectors, are transported into cells of the host plant to their targets remain a matter of speculation. Accumulating evidence from preliminary studies suggests that genetic material may be transferred from the host plant to the infecting fungal pathogen cell through exosomal biogenesis pathways [117]. At fungal penetration sites, multivesicular compartments aggregate around fungal haustorial complexes in the host cytoplasm, allowing differentiated vesicle trafficking across the plant-pathogen cellular interface to occur anterogradely, and possibly retrogradely. These multivesicular bodies consist of several intraluminal vesicles, which are discharged extracellularly as exosomes into the paramural region after fusion with the plasma membrane [117]. Multivesicular body-like compartments were reported to be involved in trafficking processes at intercellular channels known as gap junctions, nanotubes, and even the internalization of plasma membrane sections by neighboring cells [124]. siRNA species generated in the host silencing donor were suggested to be transmitted to the fungal recipient through an exocytic/endocytic exchange process at the haustorial interface [117]. Exosomes and plasma membrane-budded microvesicles have both been identified as extracellular vesicles (EVs) that are secreted by plant cells and found in the cells of fungal pathogens [125]. EVs serve as mediators of infection and defenses during plant–fungal pathogen interactions (Figure 1). Active extracellular vesicular transport, passive transport via trans-cell wall diffusion, binding and internalization through membrane-associated receptors, and other trans-membrane pores or channels are all possible sRNA trafficking mechanisms across the plant-fungal interface [81][86][126][127][128][129][130]. A diverse collection of plant sRNAs, including miRNAs and siRNAs, are selectively loaded into the EVs of plant cells [55][129]. It was suggested that fungal sRNAs delivery is facilitated by EVs, similar to the suggested plant-extracellular vesicle-mediated sRNA transport [131]. To test this hypothesis, EVs isolated from various fungal pathogens including F. oxysporum [132][133], F. graminearum [133][134], Z. tritici [135], and Ustilago maydis [136], were established and this laid a foundation for future study of cross-kingdom RNA transport in plant-fungal pathogen interactions [137]. The secreted EVs included a variety of membrane-trafficking proteins and numerous proteins for substrate transport, indicating that EVs might serve an important role in RNA trafficking [127]. Altogether, accumulating evidence points to the idea that fungal EVs are a viable method of transporting pathogen effector proteins, sRNAs and SMs into host plant cells, and their packaging in membrane-bound compartments protects them from degradation by the host enzymes and dilution by water in the plant apoplast.

7. Tools for the Prediction and Study of SM and sRNA Effectors

In effector biology, genomes, transcriptomes, proteomes, and metabolomes are mined to facilitate the discovery of potential effector genes for molecular or cellular biology, biochemistry, and reverse genetics. Future studies need to focus on the development of integrated approaches for the molecular and functional characterization of fungal SM and sRNA effectors during interactions with their host plants. Deletion mutants are usually studied for pathogenicity or symbiosis. The most significant obstacles in the generation of deletion mutants in fungi continue to be transformation and homologous recombination. This strategy has various difficulties when used for SM and sRNA effectors; therefore, new experimental approaches are needed to overcome them. Phylogenetics and comparative genomics analyses should be performed before experimental research because they are particularly informative since the number of fungal genomes and documented SM pathways is increasing. In silico studies may provide novel insights into the organization of conserved gene clusters, as well as their limits and evolutionary history. These kinds of approaches are extremely useful in locating gene clusters that play a role in the production of SMs that have been characterized in other fungal species. They also make it possible to predict the production of compounds that are either identical or related to those produced by specific fungal species. The increased availability of fungal genome sequences and next-generation genomic technologies enables the assessment of SM gene clusters in an individual fungus. RNA-Seq has revolutionized transcriptome profiling and is utilized to study SM gene cluster expression during infection. RNA-Seq can simultaneously quantify transcripts from many organisms, making it ideal for studying plant-pathogen interactions. Manipulating strain-unique SM genes involved in host-specific pathogenicity facilitates plant-fungal pathogen interactions research. Using BGC expression in heterologous hosts such as Saccharomyces cerevisiae or Aspergillus spp. may help to overcome functional redundancy and in-planta detection limitations. SMs from plant pathogenic fungi have primarily been evaluated using phytotoxicity tests. The utilization of chemical genetic screenings may also discover actions against phytohormone signaling pathways and PTI responses. For high-throughput chemical screening, this technique may utilize A. thaliana transgenic lines that express reporter genes in 96-well microplates [138]. Patterns of gene expression and regulation may be used to decipher the complicated bidirectional interaction between pathogen and host cells [90]. Using RNA-seq on both the pathogen and the host is an effective way to examine both sides of this relationship [139][140]. The recently discovered CRISPR Cas13 system can be used to study sRNA effectors, for instance, through the inactivation or localization of fungal sRNAs [141]. Recent studies have shown that extracellular vesicles play significant roles in host defense and pathogen virulence as well as being essential tools for communication between plants and pathogens. To induce the silencing of fungal genes essential for pathogenicity, plant cells secrete extracellular vesicles containing sRNAs into fungal cells. Transmission electron microscopy following ultra-rapid cryofixation showed EVs in Golovinomyces orontii extrahaustorial matrix. EVs produced by apoplastic pathogens may be detected from plant washing solutions [142]. Such fluids likely include plant and fungal EVs, making it difficult to determine their source of origin. To find out if plant pathogen EVs carry RNAs that are functional inside host plant cells, more research must be done, including the biogenesis of EVs and how specific molecules are sorted and directed towards them. 

References

  1. Rangel, L.I.; Bolton, M.D. The unsung roles of microbial secondary metabolite effectors in the plant disease cacophony. Curr. Opin. Plant Biol. 2022, 68, 102233. https://doi.org/10.1016/j.pbi.2022.102233.
  2. Jaswal, R.; Kiran, K.; Rajarammohan, S.; Dubey, H.; Singh, P.K.; Sharma, Y.; Deshmukh, R.; Sonah, H.; Gupta, N.; Sharma, T. Effector Biology of Biotrophic Plant Fungal Pathogens: Current Advances and Future Prospects. Microbiol. Res. 2020, 241, 126567. https://doi.org/10.1016/j.micres.2020.126567.
  3. Mapuranga, J.; Zhang, N.; Zhang, L.; Chang, J.; Yang, W. Infection Strategies and Pathogenicity of Biotrophic Plant Fungal Pathogens. Front. Microbiol. 2022, 13, 799396. https://doi.org/10.3389/fmicb.2022.799396.
  4. Varden, F.A.; De la Concepcion, J.C.; Maidment, J.H.; Banfield, M.J. Taking the stage: Effectors in the spotlight. Curr. Opin. Plant Biol. 2017, 38, 25–33. https://doi.org/10.1016/j.pbi.2017.04.013.
  5. Weiberg, A.; Jin, H. Small RNAs—The secret agents in the plant-pathogen interactions. Curr. Opin. Plant Biol. 2015, 26, 87–94. https://doi.org/10.1016/j.pbi.2015.05.033.
  6. Mueth, N.A.; Ramachandran, S.R.; Hulbert, S.H. Small RNAs from the wheat stripe rust fungus (Puccinia striiformis f.sp. tritici). BMC Genom. 2015, 16, 1–16. https://doi.org/10.1186/s12864-015-1895-4.
  7. Wang, B.; Sun, Y.; Song, N.; Zhao, M.; Liu, R.; Feng, H.; Wang, X.; Kang, Z. Puccinia striiformis f. sp. tritici mi croRNA ‐like RNA 1 (Pst ‐milR1), an important pathogenicity factor of Pst, impairs wheat resistance to Pst by suppressing the wheat pathogenesis‐related 2 gene. N. Phytol. 2017, 215, 338–350. https://doi.org/10.1111/nph.14577
  8. Kusch, S.; Frantzeskakis, L.; Thieron, H.; Panstruga, R. Small RNAs from cereal powdery mildew pathogens may target host plant genes. Fungal Biol. 2018, 122, 1050–1063. https://doi.org/10.1016/j.funbio.2018.08.008.
  9. Dubey, H.; Kiran, K.; Jaswal, R.; Jain, P.; Kayastha, A.M.; Bhardwaj, S.C.; Mondal, T.K.; Sharma, T.R. Discovery and profiling of small RNAs from Puccinia triticina by deep sequencing and identification of their potential targets in wheat. Funct. Integr. Genom. 2019, 19, 391–407. https://doi.org/10.1007/s10142-018-00652-1.
  10. Pusztahelyi, T.; Holb, I.J.; Pã³Csi, I. Secondary metabolites in fungus-plant interactions. Front. Plant Sci. 2015, 6, 573. https://doi.org/10.3389/fpls.2015.00573.
  11. Castro-Moretti, F.R.; Gentzel, I.N.; Mackey, D.; Alonso, A.P. Metabolomics as an Emerging Tool for the Study of Plant–Pathogen Interactions. Metabolites 2020, 10, 52. https://doi.org/10.3390/metabo10020052.
  12. Keller, N.P.; Turner, G.; Bennett, J.W. Fungal secondary metabolism—From biochemistry to genomics. Nat. Rev. Genet. 2005, 3, 937–947. https://doi.org/10.1038/nrmicro1286.
  13. Brakhage, A.A. Regulation of fungal secondary metabolism. Nat. Rev. Genet. 2012, 11, 21–32. https://doi.org/10.1038/nrmicro2916.
  14. Keller, N.P. Fungal secondary metabolism: Regulation, function and drug discovery. Nat. Rev. Microbiol. 2019, 17, 167–180. https://doi.org/10.1038/s41579-018-0121-1.
  15. Rangel, L.I.; Hamilton, O.; Jonge, R.; Bolton, M.D. Fungal social influencers: Secondary metabolites as a platform for shaping the plant‐associated community. Plant J. 2021, 108, 632–645. https://doi.org/10.1111/tpj.15490.
  16. Collemare, J.; Billard, A.; Böhnert, H.; Lebrun, M.-H. Biosynthesis of secondary metabolites in the rice blast fungus Magnaporthe grisea: The role of hybrid PKS-NRPS in pathogenicity. Mycol. Res. 2008, 112, 207–215. https://doi.org/10.1016/j.mycres.2007.08.003.
  17. Kai, K. Bioorganic chemistry of signaling molecules in microbial communication. J. Pestic. Sci. 2019, 44, 200–207. https://doi.org/10.1584/jpestics.j19-02.
  18. Lo Presti, L.; Lanver, D.; Schweizer, G.; Tanaka, S.; Liang, L.; Tollot, M.; Zuccaro, A.; Reissmann, S.; Kahmann, R. Fungal Effectors and Plant Susceptibility. Annu. Rev. Plant Biol. 2015, 66, 513–545. https://doi.org/10.1146/annurev-arplant-043014-114623.
  19. Collemare, J.; Lebrun, M. Fungal Secondary Metabolites: Ancient Toxins and Novel Effectors in Plant–Microbe Interactions. 2011, 377–400. https://doi.org/10.1002/9781119949138.ch15.Kwon, C.; Bednarek, P.; Schulze-Lefert, P. Secretory Pathways in Plant Immune Responses. Plant Physiol. 2008, 147, 1575–1583. https://doi.org/10.1104/pp.108.121566.
  20. Stergiopoulos, I.; Collemare, J.; Mehrabi, R.; De Wit, P.J. Phytotoxic secondary metabolites and peptides produced by plant pathogenic Dothideomycete fungi. FEMS Microbiol. Rev. 2013, 37, 67–93. https://doi.org/10.1111/j.1574-6976.2012.00349.x.Nielsen, M.E.; Feechan, A.; Böhlenius, H.; Ueda, T.; Thordal-Christensen, H. Arabidopsis ARF-GTP exchange factor, GNOM, mediates transport required for innate immunity and focal accumulation of syntaxin PEN1. Proc. Natl. Acad. Sci. USA 2012, 109, 11443–11448. https://doi.org/10.1073/pnas.1117596109.
  21. Friesen, T.L.; Faris, J.D.; Solomon, P.S.; Oliver, R.P. Host-specific toxins: Effectors of necrotrophic pathogenicity. Cell. Microbiol. 2008, 10, 1421–1428. https://doi.org/10.1111/j.1462-5822.2008.01153.x.Pestka, J.J. Deoxynivalenol: Mechanisms of action, human exposure, and toxicological relevance. Arch. Toxicol. 2010, 84, 663–679. https://doi.org/10.1007/s00204-010-0579-8.
  22. Griffiths, S.; Mesarich, C.H.; Overdijk, E.J.R.; Saccomanno, B.; de Wit, P.J.G.M.; Collemare, J. Down‐regulation of cladofulvin biosynthesis is required for biotrophic growth of Cladosporium fulvum on tomato. Mol. Plant Pathol. 2017, 19, 369–380. https://doi.org/10.1111/mpp.12527.Chen, Y.; Kistler, H.C.; Ma, Z. Fusarium graminearum Trichothecene Mycotoxins: Biosynthesis, Regulation, and Management. Annu. Rev. Phytopathol. 2019, 57, 15–39. https://doi.org/10.1146/annurev-phyto-082718-100318.
  23. Haas, H.; Eisendle, M.; Turgeon, B.G. Siderophores in Fungal Physiology and Virulence. Annu. Rev. Phytopathol. 2008, 46, 149–187. https://doi.org/10.1146/annurev.phyto.45.062806.094338.Xu, M.; Wang, Q.; Wang, G.; Zhang, X.; Liu, H.; Jiang, C. Combatting Fusarium head blight: Advances in molecular interactions between Fusarium graminearum and wheat. Phytopathol. Res. 2022, 4, 1–16. https://doi.org/10.1186/s42483-022-00142-0.
  24. Bushley, K.E.; Ripoll, D.R.; Turgeon, B.G. Module evolution and substrate specificity of fungal nonribosomal peptide synthetases involved in siderophore biosynthesis. BMC Evol. Biol. 2008, 8, 328. https://doi.org/10.1186/1471-2148-8-328.Navarre, D.A.; Wolpert, T.J. Victorin Induction of an Apoptotic/Senescence–like Response in Oats. Plant Cell 1999, 11, 237–249. https://doi.org/10.1105/tpc.11.2.237.
  25. Skellam, E. The biosynthesis of cytochalasans. Nat. Prod. Rep. 2017, 34, 1252–1263. https://doi.org/10.1039/c7np00036g.Tada, Y.; Kusaka, K.; Betsuyaku, S.; Shinogi, T.; Sakamoto, M.; Ohura, Y.; Hata, S.; Mori, T.; Tosa, Y.; Mayama, S. Victorin Triggers Programmed Cell Death and the Defense Response via Interaction with a Cell Surface Mediator. Plant Cell Physiol. 2005, 46, 1787–1798. https://doi.org/10.1093/pcp/pci193.
  26. Kwon, C.; Bednarek, P.; Schulze-Lefert, P. Secretory Pathways in Plant Immune Responses. Plant Physiol. 2008, 147, 1575–1583. https://doi.org/10.1104/pp.108.121566.Sweat, T.A.; Lorang, J.M.; Bakker, E.G.; Wolpert, T.J. Characterization of Natural and Induced Variation in the LOV1 Gene, a CC-NB-LRR Gene Conferring Victorin Sensitivity and Disease Susceptibility in Arabidopsis. Mol. Plant-Microbe Interact. 2008, 21, 7–19. https://doi.org/10.1094/mpmi-21-1-0007.
  27. Nielsen, M.E.; Feechan, A.; Böhlenius, H.; Ueda, T.; Thordal-Christensen, H. Arabidopsis ARF-GTP exchange factor, GNOM, mediates transport required for innate immunity and focal accumulation of syntaxin PEN1. Proc. Natl. Acad. Sci. USA 2012, 109, 11443–11448. https://doi.org/10.1073/pnas.1117596109.Brosch, G.; Ransom, R.; Lechner, T.; Walton, J.D.; Loidl, P. Inhibition of maize histone deacetylases by HC toxin, the host-selective toxin of Cochliobolus carbonum. Plant Cell 1995, 7, 1941–1950. https://doi.org/10.1105/tpc.7.11.1941.
  28. Pestka, J.J. Deoxynivalenol: Mechanisms of action, human exposure, and toxicological relevance. Arch. Toxicol. 2010, 84, 663–679. https://doi.org/10.1007/s00204-010-0579-8.Ransom, R.F.; Walton, J.D. Histone Hyperacetylation in Maize in Response to Treatment with HC-Toxin or Infection by the Filamentous Fungus Cochliobolus carbonum. Plant Physiol. 1997, 115, 1021–1027. https://doi.org/10.1104/pp.115.3.1021.
  29. Chen, Y.; Kistler, H.C.; Ma, Z. Fusarium graminearum Trichothecene Mycotoxins: Biosynthesis, Regulation, and Management. Annu. Rev. Phytopathol. 2019, 57, 15–39. https://doi.org/10.1146/annurev-phyto-082718-100318.Walton, J.D. Host-selective toxins: Agents of compatibility. Plant Cell 1996, 8, 1723–1733. https://doi.org/10.1105/tpc.8.10.1723.
  30. Xu, M.; Wang, Q.; Wang, G.; Zhang, X.; Liu, H.; Jiang, C. Combatting Fusarium head blight: Advances in molecular interactions between Fusarium graminearum and wheat. Phytopathol. Res. 2022, 4, 1–16. https://doi.org/10.1186/s42483-022-00142-0.Thomma, B.P.H.J. Alternariaspp.: From general saprophyte to specific parasite. Mol. Plant Pathol. 2003, 4, 225–236. https://doi.org/10.1046/j.1364-3703.2003.00173.x.
  31. Brown, N.A.; Evans, J.; Mead, A.; Hammond-Kosack, K.E. A spatial temporal analysis of the Fusarium graminearum tran-scriptome during symptomless and symptomatic wheat infection. Mol. Plant Pathol. 2017, 18, 1295–1312. https://doi.org/10.1111/mpp.12564.Spassieva, S.D.; Markham, J.E.; Hille, J. The plant disease resistance geneAsc-1prevents disruption of sphingolipid metabolism during AAL-toxin-induced programmed cell death. Plant J. 2002, 32, 561–572. https://doi.org/10.1046/j.1365-313x.2002.01444.x.
  32. Bönnighausen, J.; Schauer, N.; Schäfer, W.; Bormann, J. Metabolic profiling of wheat rachis node infection by Fusarium gra-minearum—Decoding deoxynivalenol-dependent susceptibility. N. Phytol. 2018, 221, 459–469. https://doi.org/10.1111/nph.15377.Wight, W.D.; Kim, K.-H.; Lawrence, C.B.; Walton, J.D. Biosynthesis and Role in Virulence of the Histone Deacetylase Inhibitor Depudecin from Alternaria brassicicola. Mol. Plant-Microbe Interact. 2009, 22, 1258–1267. https://doi.org/10.1094/mpmi-22-10-1258.
  33. Desmond, O.J.; Manners, J.M.; Stephens, A.E.; Maclean, D.J.; Schenk, P.M.; Gardiner, D.M.; Munn, A.L.; Kazan, K. The Fusarium mycotoxin deoxynivalenol elicits hydrogen peroxide production, programmed cell death and defence responses in wheat. Mol. Plant Pathol. 2008, 9, 435–445. https://doi.org/10.1111/j.1364-3703.2008.00475.x.Mapuranga, J.; Zhang, N.; Zhang, L.; Liu, W.; Chang, J.; Yang, W. Harnessing genetic resistance to rusts in wheat and inte-grated rust management methods to develop more durable resistant cultivars. Front. Plant Sci. 2022, 13, 951095. https://doi.org/10.3389/fpls.2022.951095.
  34. Diamond, M.; Reape, T.J.; Rocha, O.; Doyle, S.; Kacprzyk, J.; Doohan, F.; McCabe, P.F. The Fusarium Mycotoxin Deoxyniva-lenol Can Inhibit Plant Apoptosis-Like Programmed Cell Death. PLoS ONE 2013, 8, e69542. https://doi.org/10.1371/journal.pone.0069542.Chapman, E.J.; Carrington, J. Specialization and evolution of endogenous small RNA pathways. Nat. Rev. Genet. 2007, 8, 884–896. https://doi.org/10.1038/nrg2179.
  35. Navarre, D.A.; Wolpert, T.J. Victorin Induction of an Apoptotic/Senescence–like Response in Oats. Plant Cell 1999, 11, 237–249. https://doi.org/10.1105/tpc.11.2.237.Axtell, M.J. Classification and Comparison of Small RNAs from Plants. Annu. Rev. Plant Biol. 2013, 64, 137–159. https://doi.org/10.1146/annurev-arplant-050312-120043.
  36. Tada, Y.; Kusaka, K.; Betsuyaku, S.; Shinogi, T.; Sakamoto, M.; Ohura, Y.; Hata, S.; Mori, T.; Tosa, Y.; Mayama, S. Victorin Triggers Programmed Cell Death and the Defense Response via Interaction with a Cell Surface Mediator. Plant Cell Physiol. 2005, 46, 1787–1798. https://doi.org/10.1093/pcp/pci193.Cai, Q.; He, B.; Wang, S.; Fletcher, S.; Niu, D.; Mitter, N.; Birch, P.R.J.; Jin, H. Message in a Bubble: Shuttling Small RNAs and Proteins Between Cells and Interacting Organisms Using Extracellular Vesicles. Annu. Rev. Plant Biol. 2021, 72, 497–524. https://doi.org/10.1146/annurev-arplant-081720-010616.
  37. Sweat, T.A.; Lorang, J.M.; Bakker, E.G.; Wolpert, T.J. Characterization of Natural and Induced Variation in the LOV1 Gene, a CC-NB-LRR Gene Conferring Victorin Sensitivity and Disease Susceptibility in Arabidopsis. Mol. Plant-Microbe Interact. 2008, 21, 7–19. https://doi.org/10.1094/mpmi-21-1-0007.Ruiz-Ferrer, V.; Voinnet, O. Roles of Plant Small RNAs in Biotic Stress Responses. Annu. Rev. Plant Biol. 2009, 60, 485–510. https://doi.org/10.1146/annurev.arplant.043008.092111.
  38. Brosch, G.; Ransom, R.; Lechner, T.; Walton, J.D.; Loidl, P. Inhibition of maize histone deacetylases by HC toxin, the host-selective toxin of Cochliobolus carbonum. Plant Cell 1995, 7, 1941–1950. https://doi.org/10.1105/tpc.7.11.1941.Weiberg, A.; Wang, M.; Bellinger, M.; Jin, H. Small RNAs: A New Paradigm in Plant-Microbe Interactions. Annu. Rev. Phyto-pathol. 2014, 52, 495–516. https://doi.org/10.1146/annurev-phyto-102313-045933.
  39. Ransom, R.F.; Walton, J.D. Histone Hyperacetylation in Maize in Response to Treatment with HC-Toxin or Infection by the Filamentous Fungus Cochliobolus carbonum. Plant Physiol. 1997, 115, 1021–1027. https://doi.org/10.1104/pp.115.3.1021.Weiberg, A.; Wang, M.; Lin, F.-M.; Zhao, H.; Zhang, Z.; Kaloshian, I.; Huang, H.-D.; Jin, H. Fungal Small RNAs Suppress Plant Immunity by Hijacking Host RNA Interference Pathways. Science 2013, 342, 118–123. https://doi.org/10.1126/science.1239705.
  40. Walton, J.D. Host-selective toxins: Agents of compatibility. Plant Cell 1996, 8, 1723–1733. https://doi.org/10.1105/tpc.8.10.1723.Wang, M.; Weiberg, A.; Dellota, E., Jr.; Yamane, D.; Jin, H. Botrytis small RNA Bc-siR37 suppresses plant defense genes by cross-kingdom RNAi. RNA Biol. 2017, 14, 421–428. https://doi.org/10.1080/15476286.2017.1291112.
  41. Thomma, B.P.H.J. Alternariaspp.: From general saprophyte to specific parasite. Mol. Plant Pathol. 2003, 4, 225–236. https://doi.org/10.1046/j.1364-3703.2003.00173.x.Jian, J.; Liang, X. One Small RNA of Fusarium graminearum Targets and Silences CEBiP Gene in Common Wheat. Microor-ganisms 2019, 7, 425. https://doi.org/10.3390/microorganisms7100425.
  42. Spassieva, S.D.; Markham, J.E.; Hille, J. The plant disease resistance geneAsc-1prevents disruption of sphingolipid metabolism during AAL-toxin-induced programmed cell death. Plant J. 2002, 32, 561–572. https://doi.org/10.1046/j.1365-313x.2002.01444.x.Ji, H.; Mao, H.; Li, S.; Feng, T.; Zhang, Z.; Cheng, L.; Luo, S.; Borkovich, K.A.; Ouyang, S. Fol ‐milR1, a pathogenicity factor of Fusarium oxysporum, confers tomato wilt disease resistance by impairing host immune responses. N. Phytol. 2021, 232, 705–718. https://doi.org/10.1111/nph.17436.
  43. Wight, W.D.; Kim, K.-H.; Lawrence, C.B.; Walton, J.D. Biosynthesis and Role in Virulence of the Histone Deacetylase Inhibitor Depudecin from Alternaria brassicicola. Mol. Plant-Microbe Interact. 2009, 22, 1258–1267. https://doi.org/10.1094/mpmi-22-10-1258.Feng, H.; Zhang, Q.; Wang, Q.; Wang, X.; Liu, J.; Li, M.; Huang, L.; Kang, Z. Target of tae-miR408, a chemocyanin-like protein gene (TaCLP1), plays positive roles in wheat response to high-salinity, heavy cupric stress and stripe rust. Plant Mol. Biol. 2013, 83, 433–443. https://doi.org/10.1007/s11103-013-0101-9.
  44. A Friedman, M.; Aggarwal, V.; E Lester, G. Inhibition of epidermal DNA synthesis by cycloheximide and other inhibitors of protein synthesis. Res. Commun. Chem. Pathol. Pharmacol. 1975, 11, 311–318.Xu, W.; Meng, Y.; Wise, R.P. Mla‐ and Rom1 ‐mediated control of microRNA398 and chloroplast copper/zinc superoxide dismutase regulates cell death in response to the barley powdery mildew fungus. N. Phytol. 2013, 201, 1396–1412. https://doi.org/10.1111/nph.12598.
  45. Chen, S.; Qiang, S. Recent advances in tenuazonic acid as a potential herbicide. Pestic. Biochem. Physiol. 2017, 143, 252–257. https://doi.org/10.1016/j.pestbp.2017.01.003.Liu, J.; Cheng, X.; Liu, D.; Xu, W.; Wise, R.; Shen, Q.-H. The miR9863 Family Regulates Distinct Mla Alleles in Barley to At-tenuate NLR Receptor-Triggered Disease Resistance and Cell-Death Signaling. PLOS Genet. 2014, 10, e1004755. https://doi.org/10.1371/journal.pgen.1004755.
  46. Wicklow, D.T.; Jordan, A.M.; Gloer, J.B. Antifungal metabolites (monorden, monocillins I, II, III) from Colletotrichum grami-nicola, a systemic vascular pathogen of maize. Mycol. Res. 2009, 113, 1433–1442. https://doi.org/10.1016/j.mycres.2009.10.001.Zhao, Z.; Feng, Q.; Cao, X.; Zhu, Y.; Wang, H.; Chandran, V.; Fan, J.; Zhao, J.; Pu, M.; Li, Y.; et al. Osa‐miR167d facilitates infection of Magnaporthe oryzae in rice. J. Integr. Plant Biol. 2019, 62, 702–715. https://doi.org/10.1111/jipb.12816.
  47. Au, T.; Chick, W.S.; Leung, P. Initial kinetics of the inactivation of calmodulin by the fungal toxin ophiobolin A. Int. J. Biochem. Cell Biol. 2000, 32, 1173–1182. https://doi.org/10.1016/s1357-2725(00)00058-3.Zhang, L.-L.; Li, Y.; Zheng, Y.-P.; Wang, H.; Yang, X.; Chen, J.-F.; Zhou, S.-X.; Wang, L.-F.; Li, X.-P.; Ma, X.-C.; et al. Expressing a Target Mimic of miR156fhl-3p Enhances Rice Blast Disease Resistance without Yield Penalty by Improving SPL14 Expression. Front. Genet. 2020, 11, 327. https://doi.org/10.3389/fgene.2020.00327.
  48. Rivero-Cruz, J.F.; Macías, M.; Cerda-García-Rojas, C.M.; Mata, R. A New Phytotoxic Nonenolide from Phoma herbarum. J. Nat. Prod. 2003, 66, 511–514. https://doi.org/10.1021/np020501t.Wang, Z.; Xia, Y.; Lin, S.; Wang, Y.; Guo, B.; Song, X.; Ding, S.; Zheng, L.; Feng, R.; Chen, S.; et al. Osa-miR164a targetsOs-NAC60and negatively regulates rice immunity against the blast fungus Magnaporthe oryzae. Plant J. 2018, 95, 584–597. https://doi.org/10.1111/tpj.13972.
  49. Dallery, J.-F.; Zimmer, M.; Halder, V.; Suliman, M.; Pigné, S.; Le Goff, G.; Gianniou, D.D.; Trougakos, I.P.; Ouazzani, J.; Gasperini, D.; et al. Inhibition of jasmonate-mediated plant defences by the fungal metabolite higginsianin B. J. Exp. Bot. 2020, 71, 2910–2921. https://doi.org/10.1093/jxb/eraa061.Wang, H.; Li, Y.; Chern, M.; Zhu, Y.; Zhang, L.-L.; Lu, J.-H.; Li, X.-P.; Dang, W.-Q.; Ma, X.-C.; Yang, Z.-R.; et al. Suppression of rice miR168 improves yield, flowering time and immunity. Nat. Plants 2021, 7, 129–136. https://doi.org/10.1038/s41477-021-00852-x.
  50. Jia, L.-J.; Tang, H.-Y.; Wang, W.-Q.; Yuan, T.-L.; Wei, W.-Q.; Pang, B.; Gong, X.-M.; Wang, S.-F.; Li, Y.-J.; Zhang, D.; et al. A linear nonribosomal octapeptide from Fusarium graminearum facilitates cell-to-cell invasion of wheat. Nat. Commun. 2019, 10, 1–20. https://doi.org/10.1038/s41467-019-08726-9.Li, Y.; Zhao, S.-L.; Li, J.-L.; Hu, X.-H.; Wang, H.; Cao, X.-L.; Xu, Y.-J.; Zhao, Z.-X.; Xiao, Z.-Y.; Yang, N.; et al. Osa-miR169 Negatively Regulates Rice Immunity against the Blast Fungus Magnaporthe oryzae. Front. Plant Sci. 2017, 8, 2. https://doi.org/10.3389/fpls.2017.00002.
  51. Tang, Z.; Tang, H.; Wang, W.; Xue, Y.; Chen, D.; Tang, W.; Liu, W. Biosynthesis of a New Fusaoctaxin Virulence Factor in Fusarium graminearum Relies on a Distinct Path to Form a Guanidinoacetyl Starter Unit Priming Nonribosomal Octapeptidyl Assembly. J. Am. Chem. Soc. 2021, 143, 19719–19730. https://doi.org/10.1021/jacs.1c07770.Zhang, X.; Bao, Y.; Shan, D.; Wang, Z.; Song, X.; Wang, Z.; Wang, J.; He, L.; Wu, L.; Zhang, Z.; et al. Magnaporthe oryzae Induces the Expression of a MicroRNA to Suppress the Immune Response in Rice. Plant Physiol. 2018, 177, 352–368. https://doi.org/10.1104/pp.17.01665.
  52. Mapuranga, J.; Zhang, N.; Zhang, L.; Liu, W.; Chang, J.; Yang, W. Harnessing genetic resistance to rusts in wheat and inte-grated rust management methods to develop more durable resistant cultivars. Front. Plant Sci. 2022, 13, 951095. https://doi.org/10.3389/fpls.2022.951095.Chandran, V.; Wang, H.; Gao, F.; Cao, X.-L.; Chen, Y.-P.; Li, G.-B.; Zhu, Y.; Yang, X.-M.; Zhang, L.-L.; Zhao, Z.; et al. miR396-OsGRFs Module Balances Growth and Rice Blast Disease-Resistance. Front. Plant Sci. 2019, 9, 1999. https://doi.org/10.3389/fpls.2018.01999.
  53. Chapman, E.J.; Carrington, J. Specialization and evolution of endogenous small RNA pathways. Nat. Rev. Genet. 2007, 8, 884–896. https://doi.org/10.1038/nrg2179.Wu, L.; Zhang, Q.; Zhou, H.; Ni, F.; Wu, X.; Qi, Y. Rice MicroRNA Effector Complexes and Targets. Plant Cell 2009, 21, 3421–3435. https://doi.org/10.1105/tpc.109.070938.
  54. Axtell, M.J. Classification and Comparison of Small RNAs from Plants. Annu. Rev. Plant Biol. 2013, 64, 137–159. https://doi.org/10.1146/annurev-arplant-050312-120043.Junhua, L.; Xuemei, Y.; Jinfeng, C.; Tingting, L.; Zijin, H.; Ying, X.; Jinlu, L.; Jiqun, Z.; Mei, P.; Hui, F.; et al. Osa-miR439 Negatively Regulates Rice Immunity Against Magnaporthe oryzae. Rice Sci. 2021, 28, 156–165. https://doi.org/10.1016/j.rsci.2021.01.005.
  55. Cai, Q.; He, B.; Wang, S.; Fletcher, S.; Niu, D.; Mitter, N.; Birch, P.R.J.; Jin, H. Message in a Bubble: Shuttling Small RNAs and Proteins Between Cells and Interacting Organisms Using Extracellular Vesicles. Annu. Rev. Plant Biol. 2021, 72, 497–524. https://doi.org/10.1146/annurev-arplant-081720-010616.Xiao, Z.Y.; Wang, Q.X.; Zhao, S.L.; Wang, H.; Li, J.L.; Fan, J.; Li, Y.; Wang, W.M. MiR444b. 2 regulates resistance to Mag-naporthe oryzae and tillering in rice. Acta Phytopathol. Sin. 2017, 47, 511–522.
  56. Ruiz-Ferrer, V.; Voinnet, O. Roles of Plant Small RNAs in Biotic Stress Responses. Annu. Rev. Plant Biol. 2009, 60, 485–510. https://doi.org/10.1146/annurev.arplant.043008.092111.Qiao, L.; Zheng, L.; Sheng, C.; Zhao, H.; Jin, H.; Niu, D. Rice siR109944 suppresses plant immunity to sheath blight and impacts multiple agronomic traits by affecting auxin homeostasis. Plant J. 2020, 102, 948–964. https://doi.org/10.1111/tpj.14677.
  57. Weiberg, A.; Wang, M.; Bellinger, M.; Jin, H. Small RNAs: A New Paradigm in Plant-Microbe Interactions. Annu. Rev. Phyto-pathol. 2014, 52, 495–516. https://doi.org/10.1146/annurev-phyto-102313-045933.Katiyar-Agarwal, S.; Jin, H. Role of Small RNAs in Host-Microbe Interactions. Annu. Rev. Phytopathol. 2010, 48, 225–246. https://doi.org/10.1146/annurev-phyto-073009-114457.
  58. Weiberg, A.; Wang, M.; Lin, F.-M.; Zhao, H.; Zhang, Z.; Kaloshian, I.; Huang, H.-D.; Jin, H. Fungal Small RNAs Suppress Plant Immunity by Hijacking Host RNA Interference Pathways. Science 2013, 342, 118–123. https://doi.org/10.1126/science.1239705.Huang, C.-Y.; Wang, H.; Hu, P.; Hamby, R.; Jin, H. Small RNAs–Big Players in Plant-Microbe Interactions. Cell Host Microbe 2019, 26, 173–182. https://doi.org/10.1016/j.chom.2019.07.021.
  59. Wang, M.; Weiberg, A.; Dellota, E., Jr.; Yamane, D.; Jin, H. Botrytis small RNA Bc-siR37 suppresses plant defense genes by cross-kingdom RNAi. RNA Biol. 2017, 14, 421–428. https://doi.org/10.1080/15476286.2017.1291112.Schaefer, L.K.; Parlange, F.; Buchmann, G.; Jung, E.; Wehrli, A.; Herren, G.; Müller, M.C.; Stehlin, J.; Schmid, R.; Wicker, T.; et al. Cross-Kingdom RNAi of Pathogen Effectors Leads to Quantitative Adult Plant Resistance in Wheat. Front. Plant Sci. 2020, 11, 253. https://doi.org/10.3389/fpls.2020.00253.
  60. Jian, J.; Liang, X. One Small RNA of Fusarium graminearum Targets and Silences CEBiP Gene in Common Wheat. Microor-ganisms 2019, 7, 425. https://doi.org/10.3390/microorganisms7100425.Navarro, L.; Dunoyer, P.; Jay, F.; Arnold, B.; Dharmasiri, N.; Estelle, M.; Voinnet, O.; Jones, J.D.G. A Plant miRNA Contributes to Antibacterial Resistance by Repressing Auxin Signaling. Science 2006, 312, 436–439. https://doi.org/10.1126/science.1126088.
  61. Ji, H.; Mao, H.; Li, S.; Feng, T.; Zhang, Z.; Cheng, L.; Luo, S.; Borkovich, K.A.; Ouyang, S. Fol ‐milR1, a pathogenicity factor of Fusarium oxysporum, confers tomato wilt disease resistance by impairing host immune responses. N. Phytol. 2021, 232, 705–718. https://doi.org/10.1111/nph.17436.Li, Y.; Lu, Y.-G.; Shi, Y.; Wu, L.; Xu, Y.-J.; Huang, F.; Guo, X.-Y.; Zhang, Y.; Fan, J.; Zhao, J.-Q.; et al. Multiple Rice MicroRNAs Are Involved in Immunity against the Blast Fungus Magnaporthe oryzae. Plant Physiol. 2013, 164, 1077–1092. https://doi.org/10.1104/pp.113.230052.
  62. Feng, H.; Zhang, Q.; Wang, Q.; Wang, X.; Liu, J.; Li, M.; Huang, L.; Kang, Z. Target of tae-miR408, a chemocyanin-like protein gene (TaCLP1), plays positive roles in wheat response to high-salinity, heavy cupric stress and stripe rust. Plant Mol. Biol. 2013, 83, 433–443. https://doi.org/10.1007/s11103-013-0101-9.Knip, M.; Constantin, M.E.; Thordal-Christensen, H. Trans-kingdom Cross-Talk: Small RNAs on the Move. PLOS Genet. 2014, 10, e1004602. https://doi.org/10.1371/journal.pgen.1004602.
  63. Xu, W.; Meng, Y.; Wise, R.P. Mla‐ and Rom1 ‐mediated control of microRNA398 and chloroplast copper/zinc superoxide dismutase regulates cell death in response to the barley powdery mildew fungus. N. Phytol. 2013, 201, 1396–1412. https://doi.org/10.1111/nph.12598.Weiberg, A.; Bellinger, M.; Jin, H. Conversations between kingdoms: Small RNAs. Curr. Opin. Biotechnol. 2015, 32, 207–215. https://doi.org/10.1016/j.copbio.2014.12.025.
  64. Liu, J.; Cheng, X.; Liu, D.; Xu, W.; Wise, R.; Shen, Q.-H. The miR9863 Family Regulates Distinct Mla Alleles in Barley to At-tenuate NLR Receptor-Triggered Disease Resistance and Cell-Death Signaling. PLOS Genet. 2014, 10, e1004755. https://doi.org/10.1371/journal.pgen.1004755.Chacko, N.; Lin, X. Non-coding RNAs in the development and pathogenesis of eukaryotic microbes. Appl. Microbiol. Biotechnol. 2013, 97, 7989–7997. https://doi.org/10.1007/s00253-013-5160-y.
  65. Zhao, Z.; Feng, Q.; Cao, X.; Zhu, Y.; Wang, H.; Chandran, V.; Fan, J.; Zhao, J.; Pu, M.; Li, Y.; et al. Osa‐miR167d facilitates infection of Magnaporthe oryzae in rice. J. Integr. Plant Biol. 2019, 62, 702–715. https://doi.org/10.1111/jipb.12816.Buck, A.H.; Coakley, G.; Simbari, F.; McSorley, H.J.; Quintana, J.F.; Le Bihan, T.; Kumar, S.; Abreu-Goodger, C.; Lear, M.; Harcus, Y.; et al. Exosomes secreted by nematode parasites transfer small RNAs to mammalian cells and modulate innate immunity. Nat. Commun. 2014, 5, 5488. https://doi.org/10.1038/ncomms6488.
  66. Zhang, L.-L.; Li, Y.; Zheng, Y.-P.; Wang, H.; Yang, X.; Chen, J.-F.; Zhou, S.-X.; Wang, L.-F.; Li, X.-P.; Ma, X.-C.; et al. Expressing a Target Mimic of miR156fhl-3p Enhances Rice Blast Disease Resistance without Yield Penalty by Improving SPL14 Expression. Front. Genet. 2020, 11, 327. https://doi.org/10.3389/fgene.2020.00327.Romano, N.; Macino, G. Quelling: Transient inactivation of gene expression in Neurospora crassa by transformation with homologous sequences. Mol. Microbiol. 1992, 6, 3343–3353. https://doi.org/10.1111/j.1365-2958.1992.tb02202.x.
  67. Wang, Z.; Xia, Y.; Lin, S.; Wang, Y.; Guo, B.; Song, X.; Ding, S.; Zheng, L.; Feng, R.; Chen, S.; et al. Osa-miR164a targetsOs-NAC60and negatively regulates rice immunity against the blast fungus Magnaporthe oryzae. Plant J. 2018, 95, 584–597. https://doi.org/10.1111/tpj.13972.Koch, A.; Kogel, K.-H. New wind in the sails: Improving the agronomic value of crop plants through RNAi-mediated gene silencing. Plant Biotechnol. J. 2014, 12, 821–831. https://doi.org/10.1111/pbi.12226.
  68. Wang, H.; Li, Y.; Chern, M.; Zhu, Y.; Zhang, L.-L.; Lu, J.-H.; Li, X.-P.; Dang, W.-Q.; Ma, X.-C.; Yang, Z.-R.; et al. Suppression of rice miR168 improves yield, flowering time and immunity. Nat. Plants 2021, 7, 129–136. https://doi.org/10.1038/s41477-021-00852-x.Nunes, C.C.; Dean, R.A. Host-induced gene silencing: A tool for understanding fungal host interaction and for developing novel disease control strategies. Mol. Plant Pathol. 2011, 13, 519–529. https://doi.org/10.1111/j.1364-3703.2011.00766.x.
  69. Li, Y.; Zhao, S.-L.; Li, J.-L.; Hu, X.-H.; Wang, H.; Cao, X.-L.; Xu, Y.-J.; Zhao, Z.-X.; Xiao, Z.-Y.; Yang, N.; et al. Osa-miR169 Negatively Regulates Rice Immunity against the Blast Fungus Magnaporthe oryzae. Front. Plant Sci. 2017, 8, 2. https://doi.org/10.3389/fpls.2017.00002.Bartel, D.P. MicroRNAs: Genomics, Biogenesis, Mechanism, and Function. Cell 2004, 116, 281–297. https://doi.org/10.1016/s0092-8674(04)00045-5.
  70. Zhang, X.; Bao, Y.; Shan, D.; Wang, Z.; Song, X.; Wang, Z.; Wang, J.; He, L.; Wu, L.; Zhang, Z.; et al. Magnaporthe oryzae Induces the Expression of a MicroRNA to Suppress the Immune Response in Rice. Plant Physiol. 2018, 177, 352–368. https://doi.org/10.1104/pp.17.01665.Wang, M.; Weiberg, A.; Jin, H. Pathogen small RNAs: A new class of effectors for pathogen attacks. Mol. Plant Pathol. 2015, 16, 219–223. https://doi.org/10.1111/mpp.12233.
  71. Chandran, V.; Wang, H.; Gao, F.; Cao, X.-L.; Chen, Y.-P.; Li, G.-B.; Zhu, Y.; Yang, X.-M.; Zhang, L.-L.; Zhao, Z.; et al. miR396-OsGRFs Module Balances Growth and Rice Blast Disease-Resistance. Front. Plant Sci. 2019, 9, 1999. https://doi.org/10.3389/fpls.2018.01999.Wang, M.; Weiberg, A.; Lin, F.-M.; Thomma, B.P.H.J.; Huang, H.-D.; Jin, H. Bidirectional cross-kingdom RNAi and fungal uptake of external RNAs confer plant protection. Nat. Plants 2016, 2, 1–10. https://doi.org/10.1038/nplants.2016.151.
  72. Wu, L.; Zhang, Q.; Zhou, H.; Ni, F.; Wu, X.; Qi, Y. Rice MicroRNA Effector Complexes and Targets. Plant Cell 2009, 21, 3421–3435. https://doi.org/10.1105/tpc.109.070938.Derbyshire, M.C.; Mbengue, M.; Barascud, M.; Navaud, O.; Raffaele, S. Small RNAs from the plant pathogenic fungus Scle-rotinia sclerotiorum highlight candidate host target genes associated with quantitative disease resistance. bioRxiv 2018, 354076. https://doi.org/10.1101/354076.
  73. Junhua, L.; Xuemei, Y.; Jinfeng, C.; Tingting, L.; Zijin, H.; Ying, X.; Jinlu, L.; Jiqun, Z.; Mei, P.; Hui, F.; et al. Osa-miR439 Negatively Regulates Rice Immunity Against Magnaporthe oryzae. Rice Sci. 2021, 28, 156–165. https://doi.org/10.1016/j.rsci.2021.01.005.Rose, L.E.; Overdijk, E.J.R.; van Damme, M. Small RNA molecules and their role in plant disease. Eur. J. Plant Pathol. 2018, 154, 115–128. https://doi.org/10.1007/s10658-018-01614-w.
  74. Xiao, Z.Y.; Wang, Q.X.; Zhao, S.L.; Wang, H.; Li, J.L.; Fan, J.; Li, Y.; Wang, W.M. MiR444b. 2 regulates resistance to Mag-naporthe oryzae and tillering in rice. Acta Phytopathol. Sin. 2017, 47, 511–522.Hudzik, C.; Hou, Y.; Ma, W.; Axtell, M.J. Exchange of Small Regulatory RNAs between Plants and Their Pests. Plant Physiol. 2019, 182, 51–62. https://doi.org/10.1104/pp.19.00931.
  75. Qiao, L.; Zheng, L.; Sheng, C.; Zhao, H.; Jin, H.; Niu, D. Rice siR109944 suppresses plant immunity to sheath blight and impacts multiple agronomic traits by affecting auxin homeostasis. Plant J. 2020, 102, 948–964. https://doi.org/10.1111/tpj.14677.Mueth, N.A.; Hulbert, S.H. Small RNAs target native and cross-kingdom transcripts on both sides of the wheat stripe rust interaction. Genomics 2022, 114, 110526. https://doi.org/10.1016/j.ygeno.2022.110526.
  76. Katiyar-Agarwal, S.; Jin, H. Role of Small RNAs in Host-Microbe Interactions. Annu. Rev. Phytopathol. 2010, 48, 225–246. https://doi.org/10.1146/annurev-phyto-073009-114457.Li, Y.; Cao, X.; Zhu, Y.; Yang, X.; Zhang, K.; Xiao, Z.; Wang, H.; Zhao, J.; Zhang, L.; Li, G.; et al. Osa‐miR398b boosts H 2 O 2 production and rice blast disease‐resistance via multiple superoxide dismutases. N. Phytol. 2019, 222, 1507–1522. https://doi.org/10.1111/nph.15678.
  77. Huang, C.-Y.; Wang, H.; Hu, P.; Hamby, R.; Jin, H. Small RNAs–Big Players in Plant-Microbe Interactions. Cell Host Microbe 2019, 26, 173–182. https://doi.org/10.1016/j.chom.2019.07.021.Campo, S.; Peris‐Peris, C.; Siré, C.; Moreno, A.B.; Donaire, L.; Zytnicki, M.; Notredame, C.; Llave, C.; San Segundo, B. Identi-fication of a novel micro RNA (mi RNA) from rice that targets an alternatively spliced transcript of the N ramp6 (N atural resistance‐associated macrophage protein 6) gene involved in pathogen resistance. N. Phytol. 2013, 199, 212–227. https://doi.org/10.1111/nph.12292.
  78. Schaefer, L.K.; Parlange, F.; Buchmann, G.; Jung, E.; Wehrli, A.; Herren, G.; Müller, M.C.; Stehlin, J.; Schmid, R.; Wicker, T.; et al. Cross-Kingdom RNAi of Pathogen Effectors Leads to Quantitative Adult Plant Resistance in Wheat. Front. Plant Sci. 2020, 11, 253. https://doi.org/10.3389/fpls.2020.00253.Salvador-Guirao, R.; Hsing, Y.-I.; Segundo, B.S. The Polycistronic miR166k-166h Positively Regulates Rice Immunity via Post-transcriptional Control of EIN2. Front. Plant Sci. 2018, 9, 337. https://doi.org/10.3389/fpls.2018.00337.
  79. Navarro, L.; Dunoyer, P.; Jay, F.; Arnold, B.; Dharmasiri, N.; Estelle, M.; Voinnet, O.; Jones, J.D.G. A Plant miRNA Contributes to Antibacterial Resistance by Repressing Auxin Signaling. Science 2006, 312, 436–439. https://doi.org/10.1126/science.1126088.Koch, A.; Wassenegger, M. Host‐induced gene silencing–Mechanisms and applications. N. Phytol. 2021, 231, 54–59. https://doi.org/10.1111/nph.17364.
  80. Li, Y.; Lu, Y.-G.; Shi, Y.; Wu, L.; Xu, Y.-J.; Huang, F.; Guo, X.-Y.; Zhang, Y.; Fan, J.; Zhao, J.-Q.; et al. Multiple Rice MicroRNAs Are Involved in Immunity against the Blast Fungus Magnaporthe oryzae. Plant Physiol. 2013, 164, 1077–1092. https://doi.org/10.1104/pp.113.230052.Nowara, D.; Gay, A.; Lacomme, C.; Shaw, J.; Ridout, C.; Douchkov, D.; Hensel, G.; Kumlehn, J.; Schweizer, P. HIGS: Host-Induced Gene Silencing in the Obligate Biotrophic Fungal Pathogen Blumeria graminis. Plant Cell 2010, 22, 3130–3141. https://doi.org/10.1105/tpc.110.077040.
  81. Knip, M.; Constantin, M.E.; Thordal-Christensen, H. Trans-kingdom Cross-Talk: Small RNAs on the Move. PLOS Genet. 2014, 10, e1004602. https://doi.org/10.1371/journal.pgen.1004602.Huang, G.; Allen, R.; Davis, E.L.; Baum, T.J.; Hussey, R.S. Engineering broad root-knot resistance in transgenic plants by RNAi silencing of a conserved and essential root-knot nematode parasitism gene. Proc. Natl. Acad. Sci. USA 2006, 103, 14302–14306. https://doi.org/10.1073/pnas.0604698103.
  82. Weiberg, A.; Bellinger, M.; Jin, H. Conversations between kingdoms: Small RNAs. Curr. Opin. Biotechnol. 2015, 32, 207–215. https://doi.org/10.1016/j.copbio.2014.12.025.Panwar, V.; McCallum, B.; Bakkeren, G. Host-induced gene silencing of wheat leaf rust fungus Puccinia triticina pathogenicity genes mediated by the Barley stripe mosaic virus. Plant Mol. Biol. 2013, 81, 595–608. https://doi.org/10.1007/s11103-013-0022-7.
  83. Chacko, N.; Lin, X. Non-coding RNAs in the development and pathogenesis of eukaryotic microbes. Appl. Microbiol. Biotechnol. 2013, 97, 7989–7997. https://doi.org/10.1007/s00253-013-5160-y.Qi, T.; Guo, J.; Peng, H.; Liu, P.; Kang, Z.; Guo, J. Host-Induced Gene Silencing: A Powerful Strategy to Control Diseases of Wheat and Barley. Int. J. Mol. Sci. 2019, 20, 206. https://doi.org/10.3390/ijms20010206.
  84. Buck, A.H.; Coakley, G.; Simbari, F.; McSorley, H.J.; Quintana, J.F.; Le Bihan, T.; Kumar, S.; Abreu-Goodger, C.; Lear, M.; Harcus, Y.; et al. Exosomes secreted by nematode parasites transfer small RNAs to mammalian cells and modulate innate immunity. Nat. Commun. 2014, 5, 5488. https://doi.org/10.1038/ncomms6488.Koch, A.; Kumar, N.; Weber, L.; Keller, H.; Imani, J.; Kogel, K.-H. Host-induced gene silencing of cytochrome P450 lanosterol C14α-demethylase–encoding genes confers strong resistance to Fusarium species. Proc. Natl. Acad. Sci. USA 2013, 110, 19324–19329. https://doi.org/10.1073/pnas.1306373110.
  85. Romano, N.; Macino, G. Quelling: Transient inactivation of gene expression in Neurospora crassa by transformation with homologous sequences. Mol. Microbiol. 1992, 6, 3343–3353. https://doi.org/10.1111/j.1365-2958.1992.tb02202.x.Pliego, C.; Nowara, D.; Bonciani, G.; Gheorghe, D.M.; Xu, R.; Surana, P.; Whigham, E.; Nettleton, D.; Bogdanove, A.J.; Wise, R.P.; et al. Host-Induced Gene Silencing in Barley Powdery Mildew Reveals a Class of Ribonuclease-Like Effectors. Mol. Plant-Microbe Interact. 2013, 26, 633–642. https://doi.org/10.1094/mpmi-01-13-0005-r.
  86. Koch, A.; Kogel, K.-H. New wind in the sails: Improving the agronomic value of crop plants through RNAi-mediated gene silencing. Plant Biotechnol. J. 2014, 12, 821–831. https://doi.org/10.1111/pbi.12226.Cheng, W.; Song, X.-S.; Li, H.P.; Cao, L.-H.; Sun, K.; Qiu, X.-L.; Xu, Y.-B.; Yang, P.; Huang, T.; Zhang, J.-B.; et al. Host-induced gene silencing of an essential chitin synthase gene confers durable resistance to Fusarium head blight and seedling blight in wheat. Plant Biotechnol. J. 2015, 13, 1335–1345. https://doi.org/10.1111/pbi.12352.
  87. Nunes, C.C.; Dean, R.A. Host-induced gene silencing: A tool for understanding fungal host interaction and for developing novel disease control strategies. Mol. Plant Pathol. 2011, 13, 519–529. https://doi.org/10.1111/j.1364-3703.2011.00766.x.Chen, W.; Kastner, C.; Nowara, D.; Oliveira-Garcia, E.; Rutten, T.; Zhao, Y.; Deising, H.B.; Kumlehn, J.; Schweizer, P. Host-induced silencing of Fusarium culmorum genes protects wheat from infection. J. Exp. Bot. 2016, 67, 4979–4991. https://doi.org/10.1093/jxb/erw263.
  88. Bartel, D.P. MicroRNAs: Genomics, Biogenesis, Mechanism, and Function. Cell 2004, 116, 281–297. https://doi.org/10.1016/s0092-8674(04)00045-5.Zhu, X.; Qi, T.; Yang, Q.; He, F.; Tan, C.; Ma, W.; Voegele, R.T.; Kang, Z.; Guo, J. Host-Induced Gene Silencing of the MAPKK Gene PsFUZ7 Confers Stable Resistance to Wheat Stripe Rust. Plant Physiol. 2017, 175, 1853–1863. https://doi.org/10.1104/pp.17.01223.
  89. Wang, M.; Weiberg, A.; Jin, H. Pathogen small RNAs: A new class of effectors for pathogen attacks. Mol. Plant Pathol. 2015, 16, 219–223. https://doi.org/10.1111/mpp.12233.Panwar, V.; Jordan, M.; McCallum, B.; Bakkeren, G. Host-induced silencing of essential genes in Puccinia triticina through transgenic expression of RNAi sequences reduces severity of leaf rust infection in wheat. Plant Biotechnol. J. 2017, 16, 1013–1023. https://doi.org/10.1111/pbi.12845.
  90. Wang, M.; Weiberg, A.; Lin, F.-M.; Thomma, B.P.H.J.; Huang, H.-D.; Jin, H. Bidirectional cross-kingdom RNAi and fungal uptake of external RNAs confer plant protection. Nat. Plants 2016, 2, 1–10. https://doi.org/10.1038/nplants.2016.151.Qi, T.; Zhu, X.; Tan, C.; Liu, P.; Guo, J.; Kang, Z.; Guo, J. Host-induced gene silencing of an important pathogenicity factor PsCPK1 in Puccinia striiformis f. sp. tritici enhances resistance of wheat to stripe rust. Plant Biotechnol. J. 2017, 16, 797–807. https://doi.org/10.1111/pbi.12829.
  91. Derbyshire, M.C.; Mbengue, M.; Barascud, M.; Navaud, O.; Raffaele, S. Small RNAs from the plant pathogenic fungus Scle-rotinia sclerotiorum highlight candidate host target genes associated with quantitative disease resistance. bioRxiv 2018, 354076. https://doi.org/10.1101/354076.Ridout, C.; Skamnioti, P.; Porritt, O.; Sacristan, S.; Jones, J.; Brown, J.K. Multiple Avirulence Paralogues in Cereal Powdery Mildew Fungi May Contribute to Parasite Fitness and Defeat of Plant Resistance. Plant Cell 2006, 18, 2402–2414. https://doi.org/10.1105/tpc.106.043307.
  92. Zhao, J.-P.; Jiang, X.-L.; Zhang, B.-Y.; Su, X.-H. Involvement of microRNA-Mediated Gene Expression Regulation in the Pathological Development of Stem Canker Disease in Populus trichocarpa. PLoS ONE 2012, 7, e44968. https://doi.org/10.1371/journal.pone.0044968.Panwar, V.; McCallum, B.; Bakkeren, G. Endogenous silencing of Puccinia triticina pathogenicity genes through in plan-ta-expressed sequences leads to the suppression of rust diseases on wheat. Plant J. 2012, 73, 521–532. https://doi.org/10.1111/tpj.12047.
  93. Gupta, O.P.; Permar, V.; Koundal, V.; Singh, U.D.; Praveen, S. MicroRNA regulated defense responses in Triticum aestivum L. during Puccinia graminis f.sp. tritici infection. Mol. Biol. Rep. 2011, 39, 817–824. https://doi.org/10.1007/s11033-011-0803-5.Kusch, S.; Singh, M.; Thieron, H.; Spanu, P.D.; Panstruga, R. Site-specific analysis reveals candidate cross-kingdom small RNAs, tRNA and rRNA fragments, and signs of fungal RNA phasing in the barley-powdery mildew interaction. bioRxiv 2022, 501657. https://doi.org/10.1101/2022.07.26.501657.
  94. Xin, M.; Wang, Y.; Yao, Y.; Xie, C.; Peng, H.; Ni, Z.; Sun, Q. Diverse set of microRNAs are responsive to powdery mildew infection and heat stress in wheat (Triticum aestivum L.). BMC Plant Biol. 2010, 10, 123. https://doi.org/10.1186/1471-2229-10-123.Bai, G.; Shaner, G. Management and resistance in wheat and barley to fusarium head blight. Annu. Rev. Phytopathol. 2004, 42, 135–161. https://doi.org/10.1146/annurev.phyto.42.040803.140340.
  95. Lau, S.K.P.; Chow, W.-N.; Wong, A.Y.P.; Yeung, J.M.Y.; Bao, J.; Zhang, N.; Lok, S.; Woo, P.C.Y.; Yuen, K.-Y. Identification of MicroRNA-Like RNAs in Mycelial and Yeast Phases of the Thermal Dimorphic Fungus Penicillium marneffei. PLOS Neglected Trop. Dis. 2013, 7, e2398. https://doi.org/10.1371/journal.pntd.0002398.Scherm, B.; Balmas, V.; Spanu, F.; Pani, G.; Delogu, G.; Pasquali, M.; Migheli, Q. Fusarium culmorum: Causal agent of foot and root rot and head blight on wheat. Mol. Plant Pathol. 2012, 14, 323–341. https://doi.org/10.1111/mpp.12011.
  96. Rose, L.E.; Overdijk, E.J.R.; van Damme, M. Small RNA molecules and their role in plant disease. Eur. J. Plant Pathol. 2018, 154, 115–128. https://doi.org/10.1007/s10658-018-01614-w.Dean, R.; Van Kan, J.A.L.; Pretorius, Z.A.; Hammond-Kosack, K.E.; Di Pietro, A.; Spanu, P.D.; Rudd, J.J.; Dickman, M.; Kahmann, R.; Ellis, J.; et al. The Top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 2012, 13, 414–430. https://doi.org/10.1111/j.1364-3703.2011.00783.x.
  97. Hudzik, C.; Hou, Y.; Ma, W.; Axtell, M.J. Exchange of Small Regulatory RNAs between Plants and Their Pests. Plant Physiol. 2019, 182, 51–62. https://doi.org/10.1104/pp.19.00931.Zhu, L.; Zhu, J.; Liu, Z.; Wang, Z.; Zhou, C.; Wang, H. Host-Induced Gene Silencing of Rice Blast Fungus Magnaporthe oryzae Pathogenicity Genes Mediated by the Brome Mosaic Virus. Genes 2017, 8, 241. https://doi.org/10.3390/genes8100241.
  98. Sperschneider, J.; Jones, A.W.; Nasim, J.; Xu, B.; Jacques, S.; Zhong, C.; Upadhyaya, N.M.; Mago, R.; Hu, Y.; Figueroa, M.; et al. The stem rust fungus Puccinia graminis f. sp. tritici induces centromeric small RNAs during late infection that are associated with genome-wide DNA methylation. BMC Biol. 2021, 19, 1–25. https://doi.org/10.1186/s12915-021-01123-z.
  99. Mueth, N.A.; Hulbert, S.H. Small RNAs target native and cross-kingdom transcripts on both sides of the wheat stripe rust interaction. Genomics 2022, 114, 110526. https://doi.org/10.1016/j.ygeno.2022.110526.
  100. Li, Y.; Cao, X.; Zhu, Y.; Yang, X.; Zhang, K.; Xiao, Z.; Wang, H.; Zhao, J.; Zhang, L.; Li, G.; et al. Osa‐miR398b boosts H 2 O 2 production and rice blast disease‐resistance via multiple superoxide dismutases. N. Phytol. 2019, 222, 1507–1522. https://doi.org/10.1111/nph.15678.
  101. Campo, S.; Peris‐Peris, C.; Siré, C.; Moreno, A.B.; Donaire, L.; Zytnicki, M.; Notredame, C.; Llave, C.; San Segundo, B. Identi-fication of a novel micro RNA (mi RNA) from rice that targets an alternatively spliced transcript of the N ramp6 (N atural resistance‐associated macrophage protein 6) gene involved in pathogen resistance. N. Phytol. 2013, 199, 212–227. https://doi.org/10.1111/nph.12292.
  102. Salvador-Guirao, R.; Hsing, Y.-I.; Segundo, B.S. The Polycistronic miR166k-166h Positively Regulates Rice Immunity via Post-transcriptional Control of EIN2. Front. Plant Sci. 2018, 9, 337. https://doi.org/10.3389/fpls.2018.00337.
  103. Zhu, C.; Liu, J.-H.; Zhao, J.-H.; Liu, T.; Chen, Y.-Y.; Wang, C.-H.; Zhang, Z.-H.; Guo, H.-S.; Duan, C.-G. A fungal effector suppresses the nuclear export of AGO1–miRNA complex to promote infection in plants. Proc. Natl. Acad. Sci. USA 2022, 119, e2114583119. https://doi.org/10.1073/pnas.2114583119.
  104. Koch, A.; Wassenegger, M. Host‐induced gene silencing–Mechanisms and applications. N. Phytol. 2021, 231, 54–59. https://doi.org/10.1111/nph.17364.
  105. Nowara, D.; Gay, A.; Lacomme, C.; Shaw, J.; Ridout, C.; Douchkov, D.; Hensel, G.; Kumlehn, J.; Schweizer, P. HIGS: Host-Induced Gene Silencing in the Obligate Biotrophic Fungal Pathogen Blumeria graminis. Plant Cell 2010, 22, 3130–3141. https://doi.org/10.1105/tpc.110.077040.
  106. Huang, G.; Allen, R.; Davis, E.L.; Baum, T.J.; Hussey, R.S. Engineering broad root-knot resistance in transgenic plants by RNAi silencing of a conserved and essential root-knot nematode parasitism gene. Proc. Natl. Acad. Sci. USA 2006, 103, 14302–14306. https://doi.org/10.1073/pnas.0604698103.
  107. Panwar, V.; McCallum, B.; Bakkeren, G. Host-induced gene silencing of wheat leaf rust fungus Puccinia triticina pathogenicity genes mediated by the Barley stripe mosaic virus. Plant Mol. Biol. 2013, 81, 595–608. https://doi.org/10.1007/s11103-013-0022-7.
  108. Qi, T.; Guo, J.; Peng, H.; Liu, P.; Kang, Z.; Guo, J. Host-Induced Gene Silencing: A Powerful Strategy to Control Diseases of Wheat and Barley. Int. J. Mol. Sci. 2019, 20, 206. https://doi.org/10.3390/ijms20010206.
  109. Koch, A.; Kumar, N.; Weber, L.; Keller, H.; Imani, J.; Kogel, K.-H. Host-induced gene silencing of cytochrome P450 lanosterol C14α-demethylase–encoding genes confers strong resistance to Fusarium species. Proc. Natl. Acad. Sci. USA 2013, 110, 19324–19329. https://doi.org/10.1073/pnas.1306373110.
  110. Pliego, C.; Nowara, D.; Bonciani, G.; Gheorghe, D.M.; Xu, R.; Surana, P.; Whigham, E.; Nettleton, D.; Bogdanove, A.J.; Wise, R.P.; et al. Host-Induced Gene Silencing in Barley Powdery Mildew Reveals a Class of Ribonuclease-Like Effectors. Mol. Plant-Microbe Interact. 2013, 26, 633–642. https://doi.org/10.1094/mpmi-01-13-0005-r.
  111. Cheng, W.; Song, X.-S.; Li, H.P.; Cao, L.-H.; Sun, K.; Qiu, X.-L.; Xu, Y.-B.; Yang, P.; Huang, T.; Zhang, J.-B.; et al. Host-induced gene silencing of an essential chitin synthase gene confers durable resistance to Fusarium head blight and seedling blight in wheat. Plant Biotechnol. J. 2015, 13, 1335–1345. https://doi.org/10.1111/pbi.12352.
  112. Chen, W.; Kastner, C.; Nowara, D.; Oliveira-Garcia, E.; Rutten, T.; Zhao, Y.; Deising, H.B.; Kumlehn, J.; Schweizer, P. Host-induced silencing of Fusarium culmorum genes protects wheat from infection. J. Exp. Bot. 2016, 67, 4979–4991. https://doi.org/10.1093/jxb/erw263.
  113. Zhu, X.; Qi, T.; Yang, Q.; He, F.; Tan, C.; Ma, W.; Voegele, R.T.; Kang, Z.; Guo, J. Host-Induced Gene Silencing of the MAPKK Gene PsFUZ7 Confers Stable Resistance to Wheat Stripe Rust. Plant Physiol. 2017, 175, 1853–1863. https://doi.org/10.1104/pp.17.01223.
  114. Panwar, V.; Jordan, M.; McCallum, B.; Bakkeren, G. Host-induced silencing of essential genes in Puccinia triticina through transgenic expression of RNAi sequences reduces severity of leaf rust infection in wheat. Plant Biotechnol. J. 2017, 16, 1013–1023. https://doi.org/10.1111/pbi.12845.
  115. Qi, T.; Zhu, X.; Tan, C.; Liu, P.; Guo, J.; Kang, Z.; Guo, J. Host-induced gene silencing of an important pathogenicity factor PsCPK1 in Puccinia striiformis f. sp. tritici enhances resistance of wheat to stripe rust. Plant Biotechnol. J. 2017, 16, 797–807. https://doi.org/10.1111/pbi.12829.
  116. Ridout, C.; Skamnioti, P.; Porritt, O.; Sacristan, S.; Jones, J.; Brown, J.K. Multiple Avirulence Paralogues in Cereal Powdery Mildew Fungi May Contribute to Parasite Fitness and Defeat of Plant Resistance. Plant Cell 2006, 18, 2402–2414. https://doi.org/10.1105/tpc.106.043307.
  117. Panwar, V.; McCallum, B.; Bakkeren, G. Endogenous silencing of Puccinia triticina pathogenicity genes through in plan-ta-expressed sequences leads to the suppression of rust diseases on wheat. Plant J. 2012, 73, 521–532. https://doi.org/10.1111/tpj.12047.
  118. Kusch, S.; Singh, M.; Thieron, H.; Spanu, P.D.; Panstruga, R. Site-specific analysis reveals candidate cross-kingdom small RNAs, tRNA and rRNA fragments, and signs of fungal RNA phasing in the barley-powdery mildew interaction. bioRxiv 2022, 501657. https://doi.org/10.1101/2022.07.26.501657.
  119. Bai, G.; Shaner, G. Management and resistance in wheat and barley to fusarium head blight. Annu. Rev. Phytopathol. 2004, 42, 135–161. https://doi.org/10.1146/annurev.phyto.42.040803.140340.
  120. Scherm, B.; Balmas, V.; Spanu, F.; Pani, G.; Delogu, G.; Pasquali, M.; Migheli, Q. Fusarium culmorum: Causal agent of foot and root rot and head blight on wheat. Mol. Plant Pathol. 2012, 14, 323–341. https://doi.org/10.1111/mpp.12011.
  121. Dean, R.; Van Kan, J.A.L.; Pretorius, Z.A.; Hammond-Kosack, K.E.; Di Pietro, A.; Spanu, P.D.; Rudd, J.J.; Dickman, M.; Kahmann, R.; Ellis, J.; et al. The Top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 2012, 13, 414–430. https://doi.org/10.1111/j.1364-3703.2011.00783.x.
  122. Zhu, L.; Zhu, J.; Liu, Z.; Wang, Z.; Zhou, C.; Wang, H. Host-Induced Gene Silencing of Rice Blast Fungus Magnaporthe oryzae Pathogenicity Genes Mediated by the Brome Mosaic Virus. Genes 2017, 8, 241. https://doi.org/10.3390/genes8100241.
  123. Guo, X.-Y.; Li, Y.; Fan, J.; Xiong, H.; Xu, F.-X.; Shi, J.; Shi, Y.; Zhao, J.-Q.; Wang, Y.-F.; Cao, X.-L.; et al. Host-Induced Gene Silencing of MoAP1 Confers Broad-Spectrum Resistance to Magnaporthe oryzae. Front. Plant Sci. 2019, 10, 433. https://doi.org/10.3389/fpls.2019.00433.
  124. Gibbings, D.; Voinnet, O. Control of RNA silencing and localization by endolysosomes. Trends Cell Biol. 2010, 20, 491–501. https://doi.org/10.1016/j.tcb.2010.06.001.
  125. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. https://doi.org/10.1038/nrm.2017.125.
  126. Chaloner, T.; van Kan, J.; Grant-Downton, R.T. RNA ‘Information Warfare’ in Pathogenic and Mutualistic Interactions. Trends Plant Sci. 2016, 21, 738–748. https://doi.org/10.1016/j.tplants.2016.05.008.
  127. Hua, C.; Zhao, J.-H.; Guo, H.-S. Trans-Kingdom RNA Silencing in Plant–Fungal Pathogen Interactions. Mol. Plant 2018, 11, 235–244. https://doi.org/10.1016/j.molp.2017.12.001.
  128. Cai, Q.; He, B.; Weiberg, A.; Buck, A.H.; Jin, H. Small RNAs and extracellular vesicles: New mechanisms of cross-species communication and innovative tools for disease control. PLOS Pathog. 2019, 15, e1008090. https://doi.org/10.1371/journal.ppat.1008090.
  129. Da Rocha, I.F.M.; Amatuzzi, R.F.; Lucena, A.C.R.; Faoro, H.; Alves, L.R. Cross-Kingdom Extracellular Vesicles EV-RNA Communication as a Mechanism for Host–Pathogen Interaction. Front. Cell. Infect. Microbiol. 2020, 10, 593160. https://doi.org/10.3389/fcimb.2020.593160.
  130. Liu, G.; Kang, G.; Wang, S.; Huang, Y.; Cai, Q. Extracellular Vesicles: Emerging Players in Plant Defense against Pathogens. Front. Plant Sci. 2021, 12, 757925. https://doi.org/10.3389/fpls.2021.757925.
  131. Kwon, S.; Tisserant, C.; Tulinski, M.; Weiberg, A.; Feldbrügge, M. Inside-out: From endosomes to extracellular vesicles in fungal RNA transport. Fungal Biol. Rev. 2020, 34, 89–99. https://doi.org/10.1016/j.fbr.2020.01.001.
  132. Bleackley, M.R.; Samuel, M.; Garcia-Ceron, D.; McKenna, J.A.; Lowe, R.; Pathan, M.; Zhao, K.; Ang, C.-S.; Mathivanan, S.; Anderson, M.A. Extracellular Vesicles from the Cotton Pathogen Fusarium oxysporum f. sp. vasinfectum Induce a Phytotoxic Response in Plants. Front. Plant Sci. 2020, 10, 1610. https://doi.org/10.3389/fpls.2019.01610.
  133. Garcia‐Ceron, D.; Dawson, C.S.; Faou, P.; Bleackley, M.R.; Anderson, M.A. Size‐exclusion chromatography allows the isolation of EVs from the filamentous fungal plant pathogen Fusarium oxysporum f. sp. vasinfectum (Fov). Proteomics 2021, 21, e2000240. https://doi.org/10.1002/pmic.202000240.
  134. Koch, A.; Schlemmer, T.; Lischka, R. Elucidating the role of extracellular vesicles in the Barley-Fusarium interaction. Trillium Extracell. Vesicles 2020, 2, 28–35. https://doi.org/10.47184/tev.2020.01.03.
  135. Hill, E.H.; Solomon, P.S. Extracellular vesicles from the apoplastic fungal wheat pathogen Zymoseptoria tritici. Fungal Biol. Biotechnol. 2020, 7, 1–14. https://doi.org/10.1186/s40694-020-00103-2.
  136. Kwon, S.; Rupp, O.; Brachmann, A.; Blum, C.; Kraege, A.; Goesmann, A.; Feldbrügge, M. mRNA Inventory of Extracellular Vesicles from Ustilago maydis. J. Fungi 2021, 7, 562. https://doi.org/10.3390/jof7070562.
  137. Schlemmer, T.; Lischka, R.; Wegner, L.; Ehlers, K.; Biedenkopf, D.; Koch, A. Extracellular vesicles isolated from dsRNA-sprayed barley plants exhibit no growth inhibition or gene silencing in Fusarium graminearum. Fungal Biol. Biotechnol. 2022, 9, 1–14. https://doi.org/10.1186/s40694-022-00143-w.
  138. Halder, V.; Kombrink, E. Facile high-throughput forward chemical genetic screening by in situ monitoring of glucuroni-dase-based reporter gene expression in Arabidopsis thaliana. Front. Plant Sci. 2015, 6, 13. https://doi.org/10.3389/fpls.2015.00013.
  139. Westermann, A.; Gorski, S.; Vogel, J. Dual RNA-seq of pathogen and host. Nat. Rev. Genet. 2012, 10, 618–630. https://doi.org/10.1038/nrmicro2852.
  140. Asai, S.; Rallapalli, G.; Piquerez, S.J.M.; Caillaud, M.-C.; Furzer, O.; Ishaque, N.; Wirthmueller, L.; Fabro, G.; Shirasu, K.; Jones, J.D.G. Expression Profiling during Arabidopsis/Downy Mildew Interaction Reveals a Highly-Expressed Effector That Attenu-ates Responses to Salicylic Acid. PLOS Pathog. 2014, 10, e1004443. https://doi.org/10.1371/journal.ppat.1004443.
  141. Murugan, K.; Babu, K.; Sundaresan, R.; Rajan, R.; Sashital, D.G. The Revolution Continues: Newly Discovered Systems Expand the CRISPR-Cas Toolkit. Mol. Cell 2017, 68, 15–25. https://doi.org/10.1016/j.molcel.2017.09.007.
  142. Rutter, B.D.; Innes, R.W. Extracellular Vesicles Isolated from the Leaf Apoplast Carry Stress-Response Proteins. Plant Physiol. 2016, 173, 728–741. https://doi.org/10.1104/pp.16.01253.
More
ScholarVision Creations