Next Article in Journal
Connecting Circuits with Networks in Addiction Neuroscience: A Salience Network Perspective
Next Article in Special Issue
Nanoparticle-Shielded dsRNA Delivery for Enhancing RNAi Efficiency in Cotton Spotted Bollworm Earias vittella (Lepidoptera: Nolidae)
Previous Article in Journal
Modulation of Membrane Trafficking of AQP5 in the Lens in Response to Changes in Zonular Tension Is Mediated by the Mechanosensitive Channel TRPV1
Previous Article in Special Issue
Role of Plant Virus Movement Proteins in Suppression of Host RNAi Defense
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNAi Technology: A New Path for the Research and Management of Obligate Biotrophic Phytopathogenic Fungi

by
Isabel Padilla-Roji
1,2,
Laura Ruiz-Jiménez
1,2,
Nisrine Bakhat
1,2,
Alejandra Vielba-Fernández
1,2,
Alejandro Pérez-García
1,2 and
Dolores Fernández-Ortuño
1,2,*
1
Departamento de Microbiología, Facultad de Ciencias, Universidad de Málaga, 29071 Málaga, Spain
2
Instituto de Hortofruticultura Subtropical y Mediterránea “La Mayora”, Universidad de Málaga, Consejo Superior de Investigaciones Científicas (IHSM-UMA-CSIC), 29071 Málaga, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(10), 9082; https://doi.org/10.3390/ijms24109082
Submission received: 5 April 2023 / Revised: 5 May 2023 / Accepted: 10 May 2023 / Published: 22 May 2023
(This article belongs to the Special Issue RNA Interference-Based Tools for Plant Improvement and Protection 2.0)

Abstract

:
Powdery mildew and rust fungi are major agricultural problems affecting many economically important crops and causing significant yield losses. These fungi are obligate biotrophic parasites that are completely dependent on their hosts for growth and reproduction. Biotrophy in these fungi is determined by the presence of haustoria, specialized fungal cells that are responsible for nutrient uptake and molecular dialogue with the host, a fact that undoubtedly complicates their study under laboratory conditions, especially in terms of genetic manipulation. RNA interference (RNAi) is the biological process of suppressing the expression of a target gene through double-stranded RNA that induces mRNA degradation. RNAi technology has revolutionized the study of these obligate biotrophic fungi by enabling the analysis of gene function in these fungal. More importantly, RNAi technology has opened new perspectives for the management of powdery mildew and rust diseases, first through the stable expression of RNAi constructs in transgenic plants and, more recently, through the non-transgenic approach called spray-induced gene silencing (SIGS). In this review, the impact of RNAi technology on the research and management of powdery mildew and rust fungi will be addressed.

Graphical Abstract

1. Introduction

RNAi is a biological mechanism in which short noncoding RNAs (sRNAs) are used to deliberately downregulate gene expression at the transcriptional or posttranscriptional level. Posttranscriptional gene silencing is a tightly controlled system that relies on a group of proteins to coordinate gene silencing based on sequence complementarity between sRNA and target mRNA [1,2]. MicroRNAs (miRNAs) and short-interfering RNAs (siRNAs) are two types of regulatory sRNAs encoded by plants. miRNAs are 20–22 nucleotide (nt) sequences formed from a single-stranded RNA molecule that folds back on itself, creating a double-stranded region with a loop called RNA hairpin (hpRNAs), whereas siRNAs are 20–24 nt sequences derived from lengthy dsRNA precursors [3,4]. RNAi regulates a variety of biological processes, including plant immunity [5], and siRNAs and miRNAs have been identified as key factors in plant defense against viruses, bacteria, and fungi [6,7,8,9]. As shown in Figure 1, the silencing process starts with the binding of a host’s ribonuclease-III called Dicer (DICER) to long dsRNAs or hpRNA and their cleavage into siRNAs of 21–25 nt in length [10,11]. DICER has a helicase domain, a Piwi/Argonaute/Zwille (PAZ) motif, a dsRNA binding domain at the N-terminus, and two RNase III motifs at the C-terminus. DICER-generated siRNAs are subsequently integrated into the RNA-induced silencing complex (RISC). This multicomponent protein complex contains an Argonaute protein (AGO) with an sRNA-binding domain and endo-nucleolytic activity for RNA cleavage, which is triggered by the ATP-dependent unwinding of the siRNA duplex [12] (Figure 1). The passenger strand is degraded, and the guide strand binds to the target mRNA sequence and stimulates endonucleolytic cleavage or inhibits translation once the siRNA is integrated into RISC [13,14]. The existence of an RNA-dependent RNA polymerase (RdRP), which can interact with the RISC complex and create new dsRNA based on the partially degraded target template utilizing the hybridized siRNA strands as primers, is assumed to be the cause of this effect (Figure 1). Then, the DICER enzyme acts on the synthetized dsRNA to make additional siRNAs (secondary siRNAs). Once a dsRNA has been delivered into a cell, its influence can last throughout development; moreover, dsRNAs can be exported to neighboring cells, spreading the knockout gene effect throughout the organism [15]. There is growing evidence that sRNAs are mobilized in bidirectional interactions between plants and their pathogens, laying the groundwork for cross-kingdom RNAi (ck-RNAi) as a plant defensive mechanism [6,9,16].
Several studies have been conducted to investigate siRNA uptake in various organisms, and two primary mechanisms for host-derived RNA absorption have been proposed. One mechanism is siRNA absorption via plant-derived extracellular vesicles (EVs), which is based on the occurrence of exosome-like vesicles in plants that can carry bioactive compounds such as sRNAs to animal cells [17,18,19] (Figure 1). For example, in the fungal pathogen Sclerotinia sclerotiorum, using live cell images, it was concluded that the uptake of dsRNA occurs via clathrin-mediated endocytosis [20]. The other proposed mechanism occurs via plasma membrane-located transporters (Figure 1). This mechanism was supported by a study with the transmembrane protein SID-1, expressed in Drosophila S2 cells, which enabled passive dsRNA uptake from a culture medium [21]. Later, the lysosome transmembrane protein SIDT2 was identified in mammals and was shown to be involved in RNA uptake and subsequent degradation in this organelle [22].
This process of RNA trafficking from plant host cells to interacting pathogens has also been described in a variety of plant pathogenic fungi and oomycetes, such as Botrytis cinerea, Cochliobolus sativus, Fusarium graminearum, Plasmopara viticola, Podosphaera xanthii, Sclerotinia sclerotiorum and Venturia inaequalis [23,24,25,26,27,28,29,30]. Currently, the mechanisms of the transfer of sRNAs from plants to pathogenic fungi are unknown; however, the discovery that these eukaryotic pathogens are inhibited by sRNAs targeting their essential and/or pathogenicity genes has raised the possibility that plants could be protected by a new generation of environmentally friendly RNA-based fungicides that can be extremely specific and easily adaptable to control multiple diseases at the same time [31]. In this review, we will address the impact of RNAi technology on the research and management of two important groups of plant fungal pathogens, powdery mildew and rust fungi. First, we provide a brief description of the biological peculiarities of these fungi. Then, we describe how RNAi approaches have contributed to the analysis of gene function and have opened up new strategies for the management of powdery mildew and rust diseases, which are among the most damaging plant diseases.

2. Powdery Mildew and Rust Fungi

Obligate biotrophic fungi are a group of the most damaging plant pathogens, incurring massive economic losses and jeopardizing global food security. Powdery mildew and rust fungi infect more than 10,000 plant species, including many agronomically important crops, such as cereals, grapevines, many vegetables, and fruits, as well as ornamental and forest plants [32]. Their complete dependence on the host to feed, grow, and reproduce significantly complicates their manipulation under laboratory conditions, hindering research on their lifestyle and pathogenicity mechanisms at the molecular level [33].
Powdery mildew fungi are phytopathogenic ascomycetes belonging to the Erysiphaceae family, order Erysiphales, which includes 900 species and more than 80 genera. They cause damage in a wide range of angiosperm hosts, including both monocotyledons and dicotyledons plants. The fungal pathogens belonging to this group are easily identified by their symptoms, including the presence of powdery white patches on leaf surfaces, petioles, stems, blooms, and even fruits [34,35] (Figure 2A). In general, powdery mildew fungi exhibit both asexual and sexual life cycles (Figure 2B). The latter is highly uncommon for some species and only occurs under suitable environmental and nutritional conditions [36]. The asexual cycle starts after a conidium settles on a susceptible host plant. After its germination, it forms a small primary germ tube that elongates to become an appressorium (Figure 2B), which is in charge of penetrating the cuticle [37]. Subsequently, a hyphal peg will penetrate the epidermal cell creating a primary haustorium [38]. Upon effective infection, the main hyphae will branch and generate secondary hyphae and secondary haustoria. Later, conidiophores will emerge vertically from hyphae, generating a varying number of conidia or asexual spores depending on the species [36,39,40,41,42] (Figure 2B). This epiphytic fungal growth causes typical powdery mildew disease signs. In the event of sexual reproduction, two mating opposite hyphae need to be in contact to create a fruiting body termed chasmothecium, which holds one or more ascus containing the ascospores or sexual spores (Figure 2B) [41,42]. Although the exact infection structures developed by ascospores have not yet been determined, it is assumed that they are similar to those developed by conidia [35,43].
On the other hand, rust fungi comprise two orders, Uredinales and Pucciniales, in the widely varied phylum of Basidiomycota formed by mushrooms and bracket fungi. Rust fungi are divided into 14 families and 166 genera. Most species are found in the genera Puccinia and Uromyces, which have approximately 5000 and 1500 taxon names listed in Index Fungorum 2013, respectively [44]. Like powdery mildews, rusts are obligate biotrophic and pathogenic fungi that live on vascular plants ranging from ferns to monocots and gymnosperms to angiosperms (Figure 3A) [45,46,47,48]. Rust fungi have a typical macrocyclic-heteroecious life cycle where meiosis occurs in short-lived basidia formed by germinating teliospores (Figure 3B). Haploid basidiospores infect the aecial host and develop protoaecia and pycnia, among other fungal structures (Figure 3(B1,B2)). Pycnial nectar droplets create haploid pycniospores and receptive hyphae, where fertilization can take place between spores, and receptive hyphae of suitable mating types (Figure 3(B2,B3)). Following plasmogamy, dikaryotic aecia differentiate inside the host, and aeciospores are liberated and distributed by the wind (Figure 3(B3)). Aeciospores infect the telial host, causing the production of uredinia and urediniospores, which is followed by recurrent cycles of vegetative development on the telial host for several weeks or months, usually throughout the summer. Uredinia develops into telia in early fall, going through an overwintering phase during which karyogamy occurs, resulting in diploid dormant teliospores Figure 3(B4,B5) [45,48,49].
Both powdery mildew and rust fungi share a special structure of parasitism developed inside plant cells termed the haustorium. This specialized cell has been shown to deploy effectors, which are secreted proteins translocated into the plant cell, responsible for promoting the manipulation of the plant’s immune system and orchestrating the reprogramming of gene expression from the infected tissue to maintain fungal growth and development upon a successful infection [50,51]. The haustorium is also involved in the uptake of nutrients such as carbohydrates and amino acids and potentially water from the host via ion pumps present in the plasma membrane [52]. In addition, its ability to take genetic material such as dsRNA or siRNA makes it a key element in the development of methods of genetic transformation for biotrophic fungi, opening a world of possibilities that will allow many processes and functions to be studied in depth in the future [53,54,55].

3. RNAi Tools for Gene Function Analysis of Obligate Biotrophic Fungi

A major limitation of molecular studies in powdery mildew and rust fungi is their genetic intractability, probably due to their lifestyle as obligate biotrophs. To date, a number of transformation methods for filamentous fungi have been developed [56]. Some have been tested in powdery mildews and rusts, but unfortunately, the transformation is unstable, and the number of transformants is very low [55,57]. To mitigate this situation in part, a number of RNAi approaches have been developed for gene function analysis of these obligate biotrophic fungi, such as virus-induced gene silencing (VIGS), host-induced gene silencing (HIGS), Agrobacterium tumefaciens-mediated HIGS (ATM-HIGS) and the direct application of dsRNA, which are described below (Figure 4):

3.1. Virus-Induced Gene Silencing (VIGS)

VIGS is a term used to describe a tool that employs recombinant viruses to induce gene silencing in response to genetically manipulated RNA viral vectors [58]; Figure 4A. This technique was described for the first time in Nicotiana benthamiana, where cDNA fragments of the N. benthamiana phytoene desaturase (PDS), a gene involved in the carotenoid biosynthesis pathway, were inserted into a hybrid viral vector composed of sequences from the tobacco and tomato mosaic viruses (TMV and ToMV). These viral constructs resulted in an inhibition of carotenoid synthesis downstream of phytoene and the rapid destruction of chlorophyll by photooxidation, resulting in a white leaf phenotype in plants [59]. Since then, this approach has become a powerful silencing tool for species where stable transformants are difficult to obtain. The most popular vector for VIGS used in monocotyledons and dicotyledons is barley stripe mosaic virus (BSMV), comprising the tripartite genome RNAa, RNAß and RNAγ [60]. RNAα encodes the replicase protein (αa), RNAβ encodes a coat protein (βa) and three movement proteins (βb, βc, and βd), and RNAγ encodes the polymerase (γa) component of the replicase and the site where the fragments of target fungal genes are inserted (usually in the antisense orientation) directly downstream of the stop codon of ORFγb. For each experimental scenario, the modified RNAγ is mixed with RNAα and RNAβ and inoculated into host plants [61]. To our knowledge, the first unique evidence of the use of VIGS for powdery mildew gene silencing was described by Nowara and colleagues, who silenced two B. graminis 1-3 β-glucosyltransferase (BgGTF1 and BgGTF2) genes using a BSMV-VIGS system and reported a reduction in fungal growth on wheat (Table 1). In rust fungi, a VIGS approach was developed to identify gene function in Puccinia striiformis f. sp. tritici. The system was used to determine the Puccinia-specific gene silencing signal from the plant to the pathogen suppressing fungal gene expression. For this proposal, five predicted secreted proteins of P. striiformis (PSTha12J12, PSTha5A23, PSTha12H2, PSTha2A5, PSTha9F18), one chitinase predicted protein (PSTha5A1) and a homologue to Uromyces fabae hexose transporter (PSTha12O3) were silenced (Table 1). While reductions in rust development or sporulation were not observed for any of the genes tested, the results showed that VIGS could be used for functional gene analyses in rust fungi [62]. Then, new rust fungal targets were studied, such as PsCNA1 and PsCNB1, which are involved in the calcineurin signaling pathway that appears to be related to rust morphogenetic haustorium differentiation during the early stage of infection and production of uredospores [63] (Table 1). Another study targeted the protein kinase gene PsSRPKL, resulting in not only a reduction in fungal growth but also an increase in reactive oxygen species (ROS) accumulation in the host [64] (Table 1). In the same species, the transient silencing of the genes encoding the adenine nucleotide translocase PsANT [65], the Zn-only superoxide dismutase PsSOD1 [66], the small GTP-binding protein PsRan [67], the MAPK kinase PsFUZ7 [68], the transcription factor PstSTE12 [69], the PKA catalytic subunit PsCPK1 [70], the MADX-box transcription factor PstMCM1-1 [71], the MAP kinase kinase kinase PsKPP4 [72], the secreted protein Pst_8713 [73], and the effector protein PstGSRE1 [74] resulted in a substantial reduction of fungal growth, diminution in the spread of the hyphae and an impaired pathogenesis capacity, with some of the genes appearing to have a role in suppressing plant immunity or cell death (Table 1). Similarly, the use of the VIGS system produced a decrease in the expression of P. triticina genes, including mitogen-activated protein kinase 1 (PtMAPK1), cyclophilin (PtCYC1) and calcineurin B (PtCNB), which are involved in the establishment of disease in host plants, reducing disease symptoms and fungal growth [75] (Table 1). Following the same methodology, other studies in P. graminis f. sp. tritici determined that the transient silencing of genes such as the putative tryptophan mono-oxygenase Pgt-IaaM or genes involved in different functions such as glycosylation, sugar metabolism, transport, thiazole biosynthesis, secreted protein or unknown function (PGTG_01136, PGTG_01215; PGTG_03478, PGTG_10731, PGTG_12890, PGTG_01304, PGTG_16914, PGTG_03590, PGTG_14350) reduces not only fungal growth but also the size of urediniospores [76,77] (Table 1).

3.2. Host Induced Gene Silencing (HIGS)

The HIGS strategy results in the silencing of a pathogen-specific gene through in planta expression of dsRNA homologous to the pathogen’s target gene of interest [15]. Micro-bombardment is one of the methods used for delivering siRNA molecules into plant cells for HIGS. The high-velocity particles penetrate the cell wall and membrane, releasing the siRNA molecules into the cell cytoplasm. Once inside the cell, siRNAs can target specific mRNA molecules of a pathogenic organism, leading to their degradation or translational repression and hence silencing the expression of the pathogen’s genes [74,78]. Control of pathogen growth occurs due to RNAi-mediated silencing of a target gene related to pathogen growth and/or development, including pathogen-related structures to pathogenesis or by silencing those that are negative regulators of the host defence. Its success is based on the ability of the powdery mildew and rust fungi to take up, presumably through the haustorium, hpRNA or other RNAi molecules produced by plant cells after transformation with the silencing constructs (Figure 4B) [75,79].
The use of the HIGS method has been mainly described for powdery mildews. It has been more than a decade since Nowara and collaborators developed the approach based on HIGS by the exchange ability of siRNA molecules between cereal cells and the obligate biotrophic fungus B. graminis f. sp. hordei through a gene silencing method using dsRNA targeting the avirulence gene Avra10. The results of this assay showed a fungal growth reduction in the absence of the resistance gene Mla10 (Table 2). These results also suggested that these fungal genes play a role in haustorium formation and elongation of secondary hyphae [53]. Subsequently, the HIGS technique was applied to study many secreted proteins in the Blumeria species, such as the silencing of eight effector candidates obtaining a significant decrease in pathogen development [54] (Table 2). Another research study analyzed the candidate-secreted effector protein (CSEP; CSEP0055), and the results showed a reduction in the formation of haustoria [80] (Table 2). Later, HIGS in other CSEPs, such as CSEP0105 and CSEP0162 [81] or CSEP0027 [82], which stabilize several intracellular factors, including defense-related signaling components or CSEP007, CSEP0025, CSEP0128, CSEP0247, CSEP0345, CSEP0420, CSEP0422, CSEP0081, and CSEP0254, which are involved in early fungal aggressiveness [83,84]. As well as CSEP0139 and CSEP0182, which suppress host cell death, also resulted in a significant reduction in fungal penetration and haustoria formation rate [85] (Table 2). Although HIGS was useful in demonstrating the role of several candidate genes, the function of many others remains unknown, leaving the door open for further research.

3.3. Agrobacterium tumefaciens-Mediated Host-Induced Gene Silencing (ATM-HIGS)

Although the so-called HIGS system has allowed the individual study of various fungal CSEPs, the method, which requires particle micro-bombardment for the transformation of plant cells, has certain disadvantages, such as low-frequency success in transformation and integration and randomness of the intracellular target (cytoplasm, nucleus, vacuole, plastid, etc.), among others [86,87]. With the finding that the virulence mechanism of Agrobacterium tumefaciens leads to tumor formation, plant biotechnologists adapted the HIGS system as a new tool for the transient transformation of plants. RNAi-based gene silencing mediated by A. tumefaciens (ATM-HIGS) uses a vector consisting of a tumor-inducing plasmid (Ti plasmid) in which the oncogenes responsible for the formation of tumors of the region known as T-DNA (transferred DNA) are replaced by the RNAi machinery for the formation of dsRNA of the target gene [88]. Thus, the Agrobacterium system produces a transient transformation in plant cells due to the delivery of RNA-silencing molecules into leaf cells or other plant tissues [89,90,91] (Figure 4C).
Panwar and collaborators performed the first Agrobacterium-mediated gene silencing a says to demonstrate its silencing ability in biotrophic fungi using genes encoding mitogen-activated protein kinase 1 (PtMAPK1), cyclophilin (PtCYC1) and calcineurin B (PtCNB) from the rust fungus P. triticina [75] (Table 3). Since then, ATM-HIGS has been used successfully for gene function analysis of candidate effectors of the powdery mildew fungus P. xanthii, such as phospholipid-binding protein (PEC019), α-mannosidase (PEC032), cellulose-binding protein (PEC054), effectors with chitinase activity (EWCAs), lytic polysaccharide mono-oxygenase (PHEC27213), chitin deacetylases (PxCDA1 and PxCDA2) and other proteins with unknown functions [55,92,93,94] (Table 3). As most of these genes contribute to fungal virulence, their knockdown resulted in a substantial restriction in fungal growth and a significant increase in the plant’s immune response (Table 3).

3.4. Direct Application of dsRNA

In recent years, the use of exogenous dsRNA, sRNAs and hpRNAs has gained prominence as a new alternative that could be regarded as more sustainable, applicable and easily introduced into the host compared to the rest of the tools already discussed [95,96] (Figure 1 and Figure 4D). In phytopathogenic fungi, there are several studies that corroborate the efficacy of the use of exogenous dsRNA molecules for several gene function analyses [23,26,97,98]. One of the examples was performed by McLoughlin and collaborators using dsRNAs directed at genes related to transcription or host colonization of the fungi Sclerotinia sclerotiorum and B. cinerea. The results of this study showed a significant decrease in fungal infection and a reduction in disease symptoms [24].
The efficacy of this technique was also evaluated by infiltration of dsRNAs targeting CSEPs in the obligate biotrophic fungus Erysiphe pisi (EpCSEP001, EpCSEP009 and EpCSP083), showing a significant reduction in disease symptoms and demonstrating the involvement of these genes in the pathogenesis of pea plants [99] (Table 4). Similarly, a functional analysis of several conserved and non-annotated proteins (CNAPs) in P. xanthii, presumably involved in essential functions such as respiration (CNAP8878, CNAP9066, CNAP10905 and CNAP30520), glycosylation (CNAP1048) and efflux transport (CNAP948), showed a potential reduction in cucurbit powdery mildew disease after the infiltration of dsRNA targeting these genes [29] (Table 4). Recently, this approach has also been tried on Asian soybean rust targeting chitin synthase (CHS) genes and resulted in a large reduction in fungal lesion formation [100] (Table 4).

4. Control of Powdery Mildew and Rust Diseases by RNAi Technology

New insights into the ability of RNA molecules to move across cellular boundaries between hosts and pathogens and their ability to specifically repress essential genes of various pathogens have led to the development of novel disease management strategies [101]. The RNAi strategies developed to control powdery mildew and rust diseases are described below.

4.1. Transgenic Plants Expressing RNAi Constructs

In the last decade, several studies have proposed the use of stable HIGS in plants to confer disease resistance to fungal pathogens [53,102,103,104]. Based on experimental validation from transient HIGS assays in barley, transgenic barley plants that expressed antifungal RNAi constructs targeting the B. graminis f. sp. hordei GTF1 gene, which encodes a 1,3-β-glucanosyltransferase belonging to the penetration-associated cap20 regulon, were tested [53,105] (Table 5). Three T1 lines showed a significant reduction in B. graminis disease symptoms when a transgenic control line, which had lost the hairpin RNAi cassette, was as susceptible to powdery mildew fungus as the non-transgenic control plants. To date, fungal pathogenesis-related genes and housekeeping genes have been the primary targets for stable HIGS; however, the increased interest in effector studies in powdery mildew encouraged the question of whether their silencing would be similar to that obtained with essential fungal genes. To answer this question, Schaefer and colleagues (2020) focused on B. graminis f. sp. tritici effectors SvrPm3a1/f1, Bgt-Bcg-6, and Bgt_Bcg-7, one of the largest classes of candidate effectors in the Blumeria genomes, belonging to the RNase-like class [106,107]. In this study, stable HIGS of the three B. graminis tritici effectors resulted in a quantitative gain of powdery mildew resistance in wheat. These resistance events could impair haustorium formation on seedlings and restrict fungal growth on leaves (Table 5).
Similarly, the expression of RNAi constructs targeting the MAPK kinase gene PsFUZ7 in transgenic wheat plants conferred strong and genetically stable resistance to the devastating stripe rust fungus P. striiformis f. sp. tritici [68] (Table 5). In this study, two independent transgenic lines, which were highly effective in restricting the spread of P. striiformis, were selected in the T3 generations and examined to verify whether this phenotype was caused by the production of siRNAs corresponding to the targeted PsFUZ7 sequences. Gene expression and biomass analyses showed that both transgenic lines exhibited a significant reduction in PsFUZ7 transcripts and fungal biomass. Moreover, histological observations revealed differential hyphal growth in transgenic lines carrying PsFUZ7 RNAi constructs compared to the control, supporting the important role of PsFUZ7 in P. striiformis virulence by regulating mycelial growth and development (Table 5). Another excellent target to generate durable genetic resistance against wheat stripe rust was PsCPK1, a protein kinase A (PKA) catalytic subunit gene from P. striiformis that is highly conserved in fungi and is involved in virulence, morphogenesis, and development [108,109]. The hairpin silencing constructs of PsCPK1 expressed in wheat plants were sufficient to suppress disease development of Pst in T4 generation lines, indicating durable resistance at the genetic level against rust infection [70] (Table 5). Generally, the T3 generation is considered the initial true transgenic line in hexaploid wheat [110]; therefore, it is significant that transgenic resistance to P. striiformis was identified up to the fourth generation [70]. In the same way, the expression of RNAi constructs targeting the Pst_4 and Pst_5 rust effectors resulted in weaker hyphal development and larger H2O2 accumulation in transgenic plants compared with the non-transgenic control plants against P. striiformis [111] (Table 5). Later, the development of transgenic wheat plants that stably expressed RNAi constructs of pathogenicity target genes of P. triticina resulted in effective resistance against wheat leaf rust (WLR) disease [112] (Table 5). In particular, the engineered resistance trait was heritable and stable in the T2 generation, and the suppression of WLR development was correlated with the presence of siRNA molecules specific to the fungal PtMAPK1 and PtCYC1 genes [75].
Although the results described previously were promising, the application of HIGS by transgenic expression may be restricted by several factors: the difficulty or impossibility of transforming several crop species, the public concern about the biosafety of genetically modified crops, and the instability of artificial RNAi constructs [111]. These factors could complicate the generation of genetically modified crops [113]; consequently, a plant disease management strategy that does not rely on transgenic approaches is highly desired for environmentally sustainable agriculture.
Table 5. RNAi transgenic plants for the control of powdery mildew and rust diseases.
Table 5. RNAi transgenic plants for the control of powdery mildew and rust diseases.
Plant HostCultivarPathogenTarget GeneGene FunctionEffectsReferences
H. vulgareGolden PromiseBlumeria graminisBgGTF11,3-β-glucanosyltransferase 1Reduced manifestation of powdery mildew symptoms[53]
T. aestivumBobwhiteB. graminis f. sp. triticiSvrPm3a1/f1RNase-like effectorEnhanced resistance to powdery mildew[9]
Bgt-Bcg-6
Bgt-Bcg-7
Xinong1376Puccinia striiformis f. sp. triticiPsFUZ7MAPK kinaseEnhanced resistance to rust[68]
PKAProtein kinase AEnhanced resistance to rust[70]
PsCPK1Catalytic subunit
FielderPst_4EffectorEnhanced resistance to rust[111]
Pst_5
FielderPuccinia triticinaPtMAPK1MAP kinaseReduction of wheat leaf rust disease symptoms[111]
PtCYC1Cyclophilin

4.2. Spray-Induced Gene Silencing (SIGS)

To circumvent transgenic approaches, an innovative new strategy designated spray-induced gene silencing (SIGS) has been recently developed, which induces the silencing of pathogen target genes without the need to develop stably transformed plants and available transformation protocols. This RNAi-based technology allows the inhibition of pathogens and disease development by topical application of siRNA or dsRNA molecules onto plants to silence essential plant pathogen genes [101] (Figure 5). To date, SIGS has been demonstrated to be effective in controlling a wide range of plant pathogenic fungi [24,26,114].
Recently, the potential of SIGS has also been tested against powdery mildew and rust diseases. The first study of suppressing cucurbit powdery mildew through SIGS was reported by Ruiz-Jiménez et al. (2021) [29]. Spray application of dsRNAs targeting three P. xanthii genes essential for fungal development induced high levels of disease control. In all cases, disease severity was reduced by approximately 80% to 90% compared to water-treated melon leaves [29] (Table 6). Furthermore, in this study, the efficacy of SIGS was tested using various doses of dsRNA, and the results indicated that such dsRNAs remained functional at concentrations as low as 5 μg/mL. However, higher concentrations of dsRNA seemed to provide higher disease control, as previously demonstrated [23]. In rust fungi, Hu et al. (2020) studied the efficacy of silencing the Asian soybean rust fungus Phakopsora pachyrhizi through SIGS. In this study, direct spraying of dsRNAs targeting genes encoding an acetyl-CoA acetyltransferase (ATC), a 40S ribosomal protein S16 (RP_S16), and a glycine cleavage system H protein (GCS_H) onto soybean leaves was able to reduce the number of pustules per cm2 of leaf, fungal biomass, and endogenous target gene expression by at least 68% compared to control soybean leaves sprayed with water [115] (Table 6). In fact, SIGS targeting P. pachyrhizi chitin synthase (CHS) genes resulted in a reduction in soybean rust lesions and appressoria formation by more than 40% [100]. On the other hand, exogenous application of dsRNA targeting essential genes of Austropuccinia psidii (the cause of myrtle rust) significantly reduced infection in whole plants [116] (Table 6).
Although further studies are needed, these early successes of SIGS approaches support the idea that RNAi technology could be used to combat powdery mildew and rust diseases in a sustainable and environmentally friendly manner. This strategy does not require the development and approval of genetically engineered technologies for each crop species. It does not limit its application to a single gene or pathogen, as it is possible to target multiple essential genes of different pathogens simultaneously [117,118]. This new class of RNA-based fungicides could offer many advantages over conventional chemical treatment. However, under field conditions, the effectiveness of dsRNAs acting as fungicides may be uncertain due to the instability of RNA molecules in the environment. For this reason, current research efforts are focusing on the use of nanoparticles as carriers to deliver biologically active dsRNA, expanding the duration of their silencing effect in field conditions [119]. Currently, the nanoparticles developed for the application of these oligonucleotides in plants include inorganic and organic nanoparticles. Among the inorganic materials, those of layered double hydroxides (LDHs), carbon dots (CDs), carbon quantum dots (CQDs) or gold nanoparticles, among others, stand out. LDH nanoparticles have been used to prolong dsRNA activity and protect against viruses [120], insects [121] and fungal pathogens [122,123]. Regarding CDs, in a recent publication, dsRNAs coated with CDs were delivered to cucumber plants, leading to promising results in the control of cucurbits viruses [124]. Regarding CQDs, Kostov and colleagues (2022) found that the mixture of CQDs with dsRNA increased RNAi efficiency by causing a significant reduction in the transcript levels of the target gene in developing sporangia [125]. On the other hand, the use of organic nanoparticles as carriers has also shown interesting results. To mimic the natural mechanisms by which plants deliver their own siRNAs to pathogens, dsRNAs packaged in liposomes or in extracellular synthetic phospholipid bilayers have been used [126]. Finally, the emergence of DNA nanotechnology has also provided a promising and highly tunable platform with which to design, synthesize and utilize DNA nanostructures to deliver cargoes (drug, DNA, RNA and protein) to bypass the plant cell wall for gene silencing applications passively. It was recently demonstrated that DNA nanostructures could be used as cargo carriers for direct siRNA delivery and gene silencing in mature tobacco plants [127].
Table 6. Control of powdery mildew and rust diseases by Spray-induced gene silencing.
Table 6. Control of powdery mildew and rust diseases by Spray-induced gene silencing.
Plant HostCultivarPathogenTarget GenePossible Gene FunctionRNA AmountRNA ApplicationEffectsReferences
C. melocv. RochetPodosphaera xanthiiPxCNAP1048Glycosylation5–30 μg/mLLeaves were spray-inoculated with 104 conidia/mL after dsRNA applicationEffective management of PM disease[29]
PxCNAP10905Respiration
PxCNAP30520
G. maxcv. EnreiPhakopsora
pachyrhizi
ATCAcetyl-CoA acyltransferase20 μg/mLLeaves were spray-inoculated with 105 uredinia/mL after dsRNA applicationEffective management of Asian soybean rust (ASR) disease[115]
RP_S1640S ribosomal protein S16
GCS_HGlycine cleavage system H protein
CHSChitin synthase10 ng/mLLeaves were drop-inoculated with 105 uredinia/mL and dsRNA simultaneouslyEffective management of Asian soybean rust (ASR) disease[100]
Syzygium jambos-Austropuccinia psidiiβ-TUBβ-tubulin100 ng/μLYoung, emerging leaves were inoculated with 1 mL of dsRNA solutionsReduction in fungal growth and in the number of urediniospores[116]
EF1-aTranslation elongation factor 1ɑ
ATCAcetyl-CoA transferase
CYP450Cytochrome P450
MAPKMitogen-activated protein kinase
GCS-HGlycine cleavage system H
28S rRNA28S ribosomal RNA
HAUS01215Haustoria target

5. Conclusions and Future Prospects

To date, chemical control has been the most effective disease management strategy against powdery mildew and rust diseases. However, the increase in public concern about the use of chemicals and the emergence of fungicide-resistant isolates have resulted in a situation where novel alternative approaches to fungicide applications are urgently needed. The discovery of cross-kingdom RNAi has provided not only new approaches for gene function studies but also a new environmentally friendly and non-transgenic tool for the management of fungal plant diseases, including those caused by powdery mildew and rust fungi. The use of RNAi-based fungicides via SIGS can circumvent the problems associated with transgenic crops through the direct application of siRNA or dsRNA molecules onto plants to provide protection against pathogens; however, the stability of these molecules under field conditions is considered a major concern that may limit the application of SIGS-based disease management strategies. Therefore, in future studies, the utilization of nanoparticles and other stabilizers could improve either dsRNA stability on plant tissues, which will reduce the application frequency for growers, or dsRNA uptake efficiency, which will reduce the amount of dsRNA needed. Another major problem to consider is the cost and low efficiency of dsRNA production. In this regard, new studies should be encouraged to develop cost-effective large-scale production of dsRNA for agricultural use to facilitate SIGS implementation. In addition, fundamental research is needed to unravel the mechanisms of sRNA uptake by these fungi, and this information may be crucial to understanding and optimizing RNAi-based gene silencing in these plant pathogens. RNAi-based fungicides and SIGS will soon be a major component of the arsenal of tools for managing powdery mildew and rust diseases, thus contributing to the advancement of the modern concept of organic and sustainable agriculture.

Author Contributions

Conceptualization, D.F.-O. and A.P.-G.; validation, N.B. and L.R.-J.; resources, I.P.-R.; writing—original draft preparation I.P.-R.; writing—review and editing, I.P.-R.; visualization, I.P.-R. and A.V.-F.; supervision, D.F.-O.; project administration, D.F.-O.; funding acquisition, D.F.-O. and A.P.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This publication is part of the R+D+i projects PID2019-107464RB-C21 and PDC2021-121373-C21, funded by MICIN/AEI//10.13039/501100011033 and by the “European Union NextGenerationEU/PRTR”. I.P.-R. was supported by a Ph.D. fellowship also from AEI, grant number PRE2020-093156.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors kindly acknowledge the excellent technical support provided by Yandira Morales and Virginia Mota.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gheysen, G.; Vanholme, B. RNAi from plants to nematodes. Trends Biotechnol. 2007, 25, 89–92. [Google Scholar] [CrossRef] [PubMed]
  2. Shabalina, S.A.; Koonin, E.V. Origins and evolution of eukaryotic RNA interference. Trends Ecol. Evol. 2008, 23, 578–587. [Google Scholar] [CrossRef] [PubMed]
  3. Borges, F.; Martienssen, R.A. The expanding world of small RNAs in plants. Nat. Rev. Mol. Cell Biol. 2015, 16, 727–741. [Google Scholar] [CrossRef] [PubMed]
  4. D’Ario, M.; Griffiths-Jones, S.; Kim, M. Small RNAs: Big impact on plant development. Trends Plant Sci. 2017, 22, 1056–1068. [Google Scholar] [CrossRef] [PubMed]
  5. Deng, Y.; Wang, J.; Tung, J.; Liu, D.; Zhou, Y.; He, S.; Du, Y.; Baker, B.; Li, F. A role for small RNA in regulating innate immunity during plant growth. PLoS Pathog. 2018, 14, e1006756. [Google Scholar] [CrossRef] [PubMed]
  6. Hua, C.; Zhao, J.H.; Guo, H.S. Trans-kingdom RNA silencing in plant–fungal pathogen interactions. Mol. Plant 2018, 11, 235–244. [Google Scholar] [CrossRef]
  7. Rosa, C.; Kuo, Y.W.; Wuriyanghan, H.; Falk, B.W. RNA interference mechanisms and applications in plant pathology. Annu. Rev. Phytopathol. 2018, 56, 581–610. [Google Scholar] [CrossRef]
  8. Muhammad, T.; Zhang, F.; Zhang, Y.; Liang, Y. RNA interference: A natural immune system of plants to counteract biotic stressors. Cells 2019, 8, 38. [Google Scholar] [CrossRef]
  9. 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. [Google Scholar] [CrossRef]
  10. Zrachya, A.; Kumar, P.P.; Ramakrishnan, U.; Levy, Y.; Loyter, A.; Arazi, T.; Lapidot, M.; Gafni, Y. Production of siRNA targeted against TYLCV coat protein transcripts leads to silencing of its expression and resistance to the virus. Transgenic Res. 2007, 16, 385–398. [Google Scholar] [CrossRef]
  11. Puyam, A.; Sharma, S.; Kashyap, P.L. RNA interference- a novel approach for plant disease management. J. Appl. Nat. Sci. 2017, 9, 1612–1618. [Google Scholar] [CrossRef]
  12. Vaucheret, H.; Vazquez, F.; Crété, P.; Bartel, D.P. The action of ARGONAUTE1 in the miRNA pathway and its regulation by the miRNA pathway are crucial for plant development. Genes Dev. 2004, 18, 1187–1197. [Google Scholar] [CrossRef] [PubMed]
  13. Limera, C.; Sabbadini, S.; Sweet, J.B.; Mezzetti, B. New biotechnological tools for the genetic improvement of major woody fruit species. Front. Plant Sci. 2017, 8, 1418. [Google Scholar] [CrossRef] [PubMed]
  14. Gebremichael, D.E.; Haile, Z.M.; Negrini, F.; Sabbadini, S.; Capriotti, L.; Mezzetti, B.; Baraldi, E. RNA interference strategies for future management of plant pathogenic fungi: Prospects and challenges. Plants 2021, 10, 650. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, D.; Rossi, J. RNAi mechanisms and applications. BioTechniques 2008, 44, 613–616. [Google Scholar] [CrossRef]
  16. Knip, M.; Constantin, M.E.; Thordal-Christensen, H. Trans-kingdom cross-talk: Small RNAs on the move. PLoS Genet. 2014, 10, e1004602. [Google Scholar] [CrossRef]
  17. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef]
  18. Mu, J.; Zhuang, X.; Wang, Q.; Jiang, H.; Deng, Z.B.; Wang, B.; Zhang, L.; Kakar, S.; Jun, Y.; Miller, D.; et al. Interspecies communication between plant and mouse gut host cells through edible plant derived exosome-like nanoparticles. Mol. Nutr. Food Res. 2014, 58, 1561–1573. [Google Scholar] [CrossRef]
  19. Raimondo, S.; Naselli, F.; Fontana, S.; Monteleone, F.; Lo Dico, A.; Saieva, L.; Zito, G.; Flugy, A.; Manno, M.; Di Bella, M.A.; et al. Citrus limon-derived nanovesicles inhibit cancer cell proliferation and suppress CML xenograft growth by inducing TRAIL-mediated cell death. Oncotarget 2015, 6, 19514–19527. [Google Scholar] [CrossRef]
  20. Wytinck, N.; Manchur, C.L.; Li, V.H.; Whyard, S.; Belmonte, M.F. dsRNA uptake in plant pests and pathogens: Insights into RNAi-based insect and fungal control technology. Plants 2020, 9, 1780. [Google Scholar] [CrossRef]
  21. Shih, J.D.; Hunter, C.P. SID-1 is a dsRNA-selective dsRNA-gated channel. RNA 2011, 17, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  22. Aizawa, S.; Fujiwara, Y.; Contu, V.R.; Hase, K.; Takahashi, M.; Kikuchi, H.; Kabuta, C.; Wada, K.; Kabuta, T. Lysosomal putative RNA transporter SIDT2 mediates direct uptake of RNA by lysosomes. Autophagy 2016, 12, 565–578. [Google Scholar] [CrossRef] [PubMed]
  23. 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, 16151. [Google Scholar] [CrossRef]
  24. McLoughlin, A.G.; Wytinck, N.; Walker, P.L.; Girard, I.J.; Rashid, K.Y.; de Kievit, T.; Fernando, W.G.D.; Whyard, S.; Belmonte, M.F. Identification and application of exogenous dsRNA confers plant protection against Sclerotinia sclerotiorum and Botrytis cinerea. Sci. Rep. 2018, 8, 7320. [Google Scholar] [CrossRef] [PubMed]
  25. Nerva, L.; Sandrini, M.; Gambino, G.; Chitarra, W. Double-stranded RNAs (dsRNAs) as a sustainable tool against gray mold (Botrytis cinerea) in grapevine: Effectiveness of different application methods in an open-air environment. Biomolecules 2020, 10, 200. [Google Scholar] [CrossRef] [PubMed]
  26. Koch, A.; Biedenkopf, D.; Furch, A.; Weber, L.; Rossbach, O.; Abdellatef, E.; Linicus, L.; Johannsmeier, J.; Jelonek, L.; Goesmann, A.; et al. An RNAi-based control of Fusarium graminearum infections through spraying of long dsRNAs involves a plant passage and is controlled by the fungal silencing machinery. PLoS Pathog 2016, 12, e1005901. [Google Scholar] [CrossRef]
  27. Werner, B.T.; Gaffar, F.Y.; Schuemann, J.; Biedenkopf, D.; Koch, A.M. RNA-spray-mediated silencing of Fusarium graminearum AGO and DCL genes improve barley disease resistance. Front. Plant Sci. 2020, 11, 476. [Google Scholar] [CrossRef]
  28. Marcianò, D.; Ricciardi, V.; Marone Fassolo, E.; Passera, A.; Bianco, P.A.; Failla, O.; Casati, P.; Maddalena, G.; De Lorenzis, G.; Toffolatti, S.L. RNAi of a putative grapevine susceptibility gene as a possible downy mildew control strategy. Front. Plant Sci. 2021, 12, 667319. [Google Scholar] [CrossRef]
  29. Ruiz-Jiménez, L.; Polonio, Á.; Vielba-Fernández, A.; Pérez-García, A.; Fernández-Ortuño, D. Gene mining for conserved, non-annotated proteins of Podosphaera xanthii identifies novel target candidates for controlling powdery mildews by spray-induced gene silencing. J Fungi 2021, 7, 735. [Google Scholar] [CrossRef]
  30. Fitzgerald, A.; van Kan, J.A.L.; Plummer, K.M. Simultaneous silencing of multiple genes in the apple scab fungus, Venturia inaequalis, by expression of RNA with chimeric inverted repeats. Fungal Genet. Biol. 2004, 41, 963–971. [Google Scholar] [CrossRef]
  31. Cai, Q.; Qiao, L.; Wang, M.; He, B.; Lin, F.M.; Palmquist, J.; Huang, S.D.; Jin, H. Plants send small RNAs in extracellular vesicles to fungal pathogen to silence virulence genes. Science 2018, 360, 1126–1129. [Google Scholar] [CrossRef]
  32. Koeck, M.; Hardham, A.R.; Dodds, P.N. The role of effectors of biotrophic and hemibiotrophic fungi in infection: Effectors of biotrophic fungi. Cell. Microbiol. 2011, 13, 1849–1857. [Google Scholar] [CrossRef]
  33. Tang, C.; Xu, Q.; Zhao, M.; Wang, X.; Kang, Z. Understanding the lifestyles and pathogenicity mechanisms of obligate biotrophic fungi in wheat: The emerging genomics era. Crop. J. 2018, 6, 60–67. [Google Scholar] [CrossRef]
  34. Braun, U. The current systematics and taxonomy of the powdery mildews (Erysiphales): An overview. Mycoscience 2011, 52, 210–212. [Google Scholar] [CrossRef]
  35. Vielba-Fernández, A.; Polonio, Á.; Ruiz-Jiménez, L.; de Vicente, A.; Pérez-García, A.; Fernández-Ortuño, D. Fungicide resistance in powdery mildew fungi. Microorganisms 2020, 8, 1431. [Google Scholar] [CrossRef]
  36. Pérez-García, A.; Romero, D.; Fernández-Ortuño, D.; López-Ruiz, F.; De Vicente, A.; Torés, J.A. The powdery mildew fungus Podosphaera fusca (synonym Podosphaera xanthii), a constant threat to cucurbits. Mol. Plant Pathol. 2009, 10, 153–160. [Google Scholar] [CrossRef]
  37. Eichmann, R.; Hückelhoven, R. Accommodation of powdery mildew fungi in intact plant cells. J. Plant Physiol. 2008, 165, 5–18. [Google Scholar] [CrossRef]
  38. Tucker, S.L.; Talbot, N.J. Surface attachment and pre-penetration stage development by plant pathogenic fungi. Annu. Rev. Phytopathol. 2001, 39, 385–417. [Google Scholar] [CrossRef]
  39. Heffer, V.; Powelson, M.L.; Johnson, K.B.; Shishkoff, N. Identification of powdery mildew fungi anno 2006. Plant Heath Instr. 2006. [Google Scholar] [CrossRef]
  40. Sidhu, G.S. Genetics of plant pathogenic fungi. In Advances in Plant Pathology; Ingram, D.S., Williams, P.H., Eds.; Academic Press: Cambridge, MA, USA, 1988; Volume 6. [Google Scholar]
  41. Gadoury, D.M.; Cadle-Davidson, L.; Wilcox, W.F.; Dry, I.B.; Seem, R.C.; Milgroom, M.G. Grapevine powdery mildew (Erysiphe necator): A fascinating system for the study of the biology, ecology and epidemiology of an obligate biotroph: Grapevine powdery mildew. Mol. Plant Pathol. 2012, 13, 1–16. [Google Scholar] [CrossRef]
  42. Saharan, G.S.; Mehta, N.K.; Meena, P.D. Infection, pathogenesis and disease cycle. In Powdery Mildew Disease of Crucifers: Biology, Ecology and Disease Management; Springer: Singapore, 2019; pp. 95–130. [Google Scholar]
  43. Jarvis, W.R.; Gubler, W.D.; Grove, G.G. Epidemiology of Powdery Mildews in Agricultural Pathosystems; Bélanger, R.R., Bushnell, W.R., Dik, A.J., Carver, T.L.W., Eds.; APS Press: St. Paul, MN, USA, 2002. [Google Scholar]
  44. Helfer, S. Rust fungi and global change. New Phytol. 2014, 201, 770–780. [Google Scholar] [CrossRef] [PubMed]
  45. Aime, M.; Toome, M.; McLaughlin, D. Pucciniomycotina. Systematics and Evolution. In The Mycota; Springer: Berlin/Heidelberg, Germany, 2014; Volume 7A. [Google Scholar]
  46. Dean, R.; Van Kan, J.A.; 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. [Google Scholar] [CrossRef] [PubMed]
  47. Fisher, M.C.; Henk, D.A.; Briggs, C.J.; Brownstein, J.S.; Madoff, L.C.; McCraw, S.L.; Gurr, S.J. Emerging fungal threats to animal, plant and ecosystem health. Nature 2012, 484, 186–194. [Google Scholar] [CrossRef] [PubMed]
  48. Lorrain, C.; Gonçalves dos Santos, K.C.; Germain, H.; Hecker, A.; Duplessis, S. Advances in understanding obligate biotrophy in rust fungi. New Phytol. 2019, 222, 1190–1206. [Google Scholar] [CrossRef] [PubMed]
  49. Hacquard, S.; Petre, B.; Frey, P.; Hecker, A.; Rouhier, N.; Duplessis, S. The poplar-poplar rust interaction: Insights from genomics and transcriptomics. J. Pathog. 2011, 2011, 716041. [Google Scholar] [CrossRef]
  50. Sohn, K.H.; Lei, R.; Nemri, A.; Jones, J.D. The downy mildew effector proteins ATR1 and ATR13 promote disease susceptibility in Arabidopsis thaliana. Plant Cell 2008, 19, 4077–4090. [Google Scholar] [CrossRef]
  51. Stergiopoulos, I.; de Wit, P.J. Fungal effector proteins. Annu. Rev. Phytopathol. 2009, 47, 233–263. [Google Scholar] [CrossRef]
  52. Polonio, A.; Pérez-García, A.; Martínez-Cruz, J.; Fernández-Ortuño, D.; de Vicente, A. The haustorium of phytopathogenic fungi: A short overview of a specialized cell of obligate biotrophic plant parasites. In Progress in Botany; Cánovas, F.M., Lüttge, U., Risueño, M.C., Pretzsch, H., Eds.; Springer Nature Switzerland AG: Basel, Switzerland, 2020; Volume 82, ISBN 978-3-030-68619-2. [Google Scholar]
  53. 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. [Google Scholar] [CrossRef]
  54. 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. [Google Scholar] [CrossRef]
  55. Martínez-Cruz, J.; Romero, D.; de la Torre, F.N.; Fernández-Ortuño, D.; Torés, J.A.; de Vicente, A.; Pérez-García, A. The functional characterization of Podosphaera xanthii candidate effector genes reveals novel target functions for fungal pathogenicity. Mol. Plant Microbe Interact. 2018, 31, 914–931. [Google Scholar] [CrossRef]
  56. Jiang, D.; Zhu, W.; Wang, Y.; Sun, C.; Zhang, K.-Q.; Yang, J. Molecular tools for functional genomics in filamentous fungi: Recent advances and new strategies. Biotech. Adv. 2013, 31, 1562–1574. [Google Scholar] [CrossRef]
  57. Martínez-Cruz, J.; Romero, D.; Vicente, A.; Pérez-García, A. Transformation of the cucurbit powdery mildew pathogen Podosphaera xanthii by Agrobacterium tumefaciens. New Phytol. 2017, 213, 1961–1973. [Google Scholar] [CrossRef] [PubMed]
  58. Becker, A.; Lange, M. VIGS—Genomics goes functional. Trends Plant Sci. 2010, 15, 1–4. [Google Scholar] [CrossRef] [PubMed]
  59. Kumagai, M.H.; Keller, Y.; Bouvier, F.; Clary, D.; Camara, B. Functional integration of non-native carotenoids into chloroplasts by viral-derived expression of capsanthin–capsorubin synthase in Nicotiana benthamiana. Plant J. 1998, 14, 305–315. [Google Scholar] [CrossRef] [PubMed]
  60. Jackson, A.O.; Lim, H.-S.; Bragg, J.; Ganesan, U.; Lee, M.Y. Hordeivirus replication, movement, and pathogenesis. Annu. Rev. Phytopathol. 2009, 47, 385–422. [Google Scholar] [CrossRef]
  61. Lee, W.-S.; Hammond-Kosack, K.E.; Kanyuka, K. Barley stripe mosaic virus—Mediated tools for investigating gene function in cereal plants and their pathogens: Virus-induced gene silencing, host-mediated gene silencing, and virus-mediated overexpression of heterologous protein. Plant Physiol. 2012, 160, 582–590. [Google Scholar] [CrossRef]
  62. Yin, C.; Jurgenson, J.E.; Hulbert, S.H. Development of a host-induced RNAi system in the wheat stripe rust fungus Puccinia striiformis f. sp. tritici. Mol. Plant Microbe Interact. 2011, 24, 554–561. [Google Scholar] [CrossRef]
  63. Zhang, H.; Guo, J.; Voegele, R.T.; Zhang, J.; Duan, Y.; Luo, H.; Kang, Z. Functional characterization of calcineurin homologs PsCNA1/PsCNB1 in Puccinia striiformis f. sp. tritici using a host-induced RNAi system. PLoS ONE 2012, 7, e49262. [Google Scholar] [CrossRef]
  64. Cheng, Y.; Wang, X.; Yao, J.; Voegele, R.T.; Zhang, Y.; Wang, W.; Huang, L.; Kang, Z. Characterization of protein kinase PsSRPKL, a novel pathogenicity factor in the wheat stripe rust fungus: Kinases and rust pathogenicity. Environ. Microbiol. 2015, 17, 2601–2617. [Google Scholar] [CrossRef]
  65. Tang, C.; Wei, J.; Han, Q.; Liu, R.; Duan, X.; Fu, Y.; Huang, X.; Wang, X.; Kang, Z. PsANT, the adenine nucleotide translocase of Puccinia striiformis, promotes cell death and fungal growth. Sci. Rep. 2015, 5, 11241. [Google Scholar] [CrossRef]
  66. Liu, J.; Guan, T.; Zheng, P.; Chen, L.; Yang, Y.; Huai, B.; Li, D.; Chang, Q.; Huang, L.; Kang, Z. An extracellular Zn-Only superoxide dismutase from Puccinia striiformis confers enhanced resistance to host-derived oxidative stress. Environ. Microbiol. 2016, 18, 4118–4135. [Google Scholar] [CrossRef] [PubMed]
  67. Cheng, Y.; Yao, J.; Zhang, Y.; Li, S.; Kang, Z. Characterization of a Ran gene from Puccinia striiformis f. sp. tritici involved in fungal growth and anti-cell death. Sci. Rep. 2016, 6, 35248. [Google Scholar] [CrossRef] [PubMed]
  68. 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. [Google Scholar] [CrossRef]
  69. Zhu, X.; Liu, W.; Chu, X.; Sun, Q.; Tan, C.; Yang, Q.; Jiao, M.; Guo, J.; Kang, Z. The transcription factor PstSTE12 is required for virulence of Puccinia striiformis f. sp. tritici. Mol. Plant Pathol. 2018, 19, 961–974. [Google Scholar] [CrossRef]
  70. 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. 2018, 16, 797–807. [Google Scholar] [CrossRef]
  71. Zhu, X.; Jiao, M.; Guo, J.; Liu, P.; Tan, C.; Yang, Q.; Zhang, Y.; Thomas Voegele, R.; Kang, Z.; Guo, J. A novel MADS-box transcription factor PstMCM1-1 is responsible for full virulence of Puccinia striiformis f. sp. tritici. Environ. Microbiol. 2018, 20, 1452–1463. [Google Scholar] [CrossRef]
  72. Zhu, X.; Guo, J.; He, F.; Zhang, Y.; Tan, C.; Yang, Q.; Huang, C.; Kang, Z.; Guo, J. Silencing PsKPP4, a MAP kinase kinase kinase gene, reduces pathogenicity of the stripe rust fungus. Mol. Plant Pathol. 2018, 19, 2590–2602. [Google Scholar] [CrossRef] [PubMed]
  73. Zhao, M.; Wang, J.; Ji, S.; Chen, Z.; Xu, J.; Tang, C.; Chen, S.; Kang, Z.; Wang, X. Candidate effector Pst_8713 impairs the plant immunity and contributes to virulence of Puccinia striiformis f. sp. tritici. Front. Plant Sci. 2018, 9, 1294. [Google Scholar] [CrossRef]
  74. Qi, T.; Guo, J.; Liu, P.; He, F.; Wan, C.; Islam, M.A.; Tyler, B.M.; Kang, Z.; Guo, J. Stripe rust effector PstGSRE1 disrupts nuclear localization of ROS-promoting transcription factor TaLOL2 to defeat ROS-induced defense in wheat. Mol. Plant 2019, 12, 1624–1638. [Google Scholar] [CrossRef]
  75. Panwar, V.; McCallum, B.; Bakkeren, G. Endogenous silencing of Puccinia triticina pathogenicity genes through in planta-expressed sequences leads to the suppression of rust diseases on wheat. Plant J. 2013, 73, 521–532. [Google Scholar] [CrossRef]
  76. Yin, C.; Park, J.-J.; Gang, D.R.; Hulbert, S.H. Characterization of a tryptophan 2-monooxygenase gene from Puccinia graminis f. sp. tritici involved in auxin biosynthesis and rust pathogenicity. Mol. Plant Microbe Interact. 2014, 27, 227–235. [Google Scholar] [CrossRef] [PubMed]
  77. Yin, C.; Downey, S.I.; Klages-Mundt, N.L.; Ramachandran, S.; Chen, X.; Szabo, L.J.; Pumphrey, M.; Hulbert, S.H. Identification of promising host-induced silencing targets among genes preferentially transcribed in haustoria of Puccinia. BMC Genomics 2015, 16, 579. [Google Scholar] [CrossRef] [PubMed]
  78. Margo, R. Genome improvement for rust disease resistance in wheat. In Genome Engineering for Crop Improvement; Sarmah, B.K., Borah, B.K., Eds.; Springer International Publishing: Cham, Switzerland, 2021; ISBN 978-3-030-63371-4. [Google Scholar]
  79. Govindarajulu, M.; Epstein, L.; Wroblewski, T.; Michelmore, R.W. Host-induced gene silencing inhibits the biotrophic pathogen causing downy mildew of lettuce. Plant Biotechnol. J. 2015, 13, 875–883. [Google Scholar] [CrossRef]
  80. Zhang, W.J.; Pedersen, C.; Kwaaitaal, M.; Gregersen, P.L.; Mørch, S.M.; Hanisch, S.; Kristensen, A.; Fuglsang, A.T.; Collinge, D.B.; Thordal-Christensen, H. Interaction of barley powdery mildew effector candidate CSEP0055 with the defence protein PR17c. Mol. Plant Pathol. 2012, 13, 1110–1119. [Google Scholar] [CrossRef]
  81. Ahmed, A.A.; Pedersen, C.; Schultz-Larsen, T.; Kwaaitaal, M.; Jørgensen, H.J.; Thordal-Christensen, H. The barley powdery mildew candidate secreted effector protein CSEP0105 inhibits the chaperone activity of a small heat shock protein. Plant Physiol. 2015, 168, 321–333. [Google Scholar] [CrossRef] [PubMed]
  82. Yuan, H.; Jin, C.; Pei, H.; Zhao, L.; Li, X.; Li, J.; Huang, W.; Fan, R.; Liu, W.; Shen, Q.H. The powdery mildew effector CSEP0027 interacts with barley catalase to regulate host immunity. Front. Plant Sci. 2021, 12, 733237. [Google Scholar] [CrossRef] [PubMed]
  83. Aguilar, G.B.; Pedersen, C.; Thordal-Christensen, H. Identification of eight effector candidate genes involved in early aggressiveness of the barley powdery mildew fungus. Plant Pathol. 2016, 65, 953–958. [Google Scholar] [CrossRef]
  84. Ahmed, A.A.; Pedersen, C.; Thordal-Christensen, H. The barley powdery mildew effector candidates CSEP0081 and CSEP0254 promote fungal infection success. PLoS ONE 2016, 11, e157586. [Google Scholar] [CrossRef]
  85. Li, X.; Jin, C.; Yuan, H.; Huang, W.; Liu, F.; Fan, R.; Xie, J.; Shen, Q.H. The barley powdery mildew effectors CSEP0139 and CSEP0182 suppress cell death and promote B. graminis fungal virulence in plants. Phytopathol. Res. 2021, 3, 7. [Google Scholar] [CrossRef]
  86. Sanford, J.C. Biolistic plant transformation. Physiol. Plant 1990, 79, 206–209. [Google Scholar] [CrossRef]
  87. Harwood, W.A. Advances and remaining challenges in the transformation of barley and wheat. J. Exp. Bot. 2012, 63, 1791–1798. [Google Scholar] [CrossRef] [PubMed]
  88. Gelvin, S.B. Agrobacterium-mediated plant transformation: The biology behind the “gene-jockeying” tool. Microbiol. Mol. Biol. Rev. 2003, 67, 16–37. [Google Scholar] [CrossRef] [PubMed]
  89. Johansen, L.K.; Carrington, J.C. Silencing on the spot. Induction and suppression of RNA silencing in the Agrobacterium-mediated transient expression system. Plant Physiol. 2001, 126, 930–938. [Google Scholar] [CrossRef] [PubMed]
  90. Duan, C.G.; Wang, C.H.; Guo, H.S. Application of RNA silencing to plant disease resistance. Silence 2012, 3, 5. [Google Scholar] [CrossRef]
  91. Bertazzon, N.; Raiola, A.; Castiglioni, C.; Gardiman, M.; Angelini, E.; Borgo, M.; Ferrari, S. Transient silencing of the grapevine gene VvPGIP1 by agroinfiltration with a construct for RNA interference. Plant Cell Rep. 2012, 31, 133–143. [Google Scholar] [CrossRef]
  92. Martínez-Cruz, J.; Romero, D.; Hierrezuelo, J.; Thon, M.; de Vicente, A.; Pérez-García, A. Effectors with chitinase activity (EWCAs), a family of conserved, secreted fungal chitinases that suppress chitin-triggered immunity. Plant Cell 2021, 33, 1319–1340. [Google Scholar] [CrossRef]
  93. Polonio, Á.; Fernández-Ortuño, D.; Vicente, A.; Pérez-García, A. A haustorial-expressed lytic polysaccharide monooxygenase from the cucurbit powdery mildew pathogen Podosphaera xanthii contributes to the suppression of chitin-triggered immunity. Mol. Plant. Pathol. 2021, 22, 580–601. [Google Scholar] [CrossRef]
  94. Martínez-Cruz, J.M.; Polonio, Á.; Zanni, R.; Romero, D.; Gálvez, J.; Fernández-Ortuño, D.; Pérez-García, A. chitin deacetylase, a novel target for the design of agricultural fungicides. J. Fungi 2021, 7, 1009. [Google Scholar] [CrossRef]
  95. Dubrovina, A.S.; Kiselev, K.V. Exogenous RNAs for gene regulation and plant resistance. Int. J. Mol. Sci. 2019, 20, 2282. [Google Scholar] [CrossRef]
  96. Dalakouras, A.; Wassenegger, M.; Dadami, E.; Ganopoulos, I.; Pappas, M.L.; Papadopoulou, K. Genetically modified organism-free RNA interference: Exogenous application of RNA molecules in plants. Plant Physiol. 2020, 182, 38–50. [Google Scholar] [CrossRef]
  97. Mumbanza, F.M.; Kiggundu, A.; Tusiime, G.; Tushemereirwe, W.K.; Niblett, C.; Bailey, A. In vitro antifungal activity of synthetic dsRNA molecules against two pathogens of banana, Fusarium oxysporum f. sp. cubense and Mycosphaerella fijiensis. Pest. Manag. Sci. 2013, 69, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, M.; Weiberg, A.; Dellota, E.; Yamane, D.; Jin, H. Botrytis small RNA Bc -siR37 suppresses plant defense genes by cross-kingdom RNAi. RNA Biol. 2017, 14, 421–428. [Google Scholar] [CrossRef] [PubMed]
  99. Sharma, G.; Aminedi, R.; Saxena, D.; Gupta, A.; Banerjee, P.; Jain, D.; Chandran, D. Effector mining from the Erysiphe pisi haustorial transcriptome identifies novel candidates involved in pea powdery mildew pathogenesis. Mol. Plant Pathol. 2019, 20, 1506–1522. [Google Scholar] [CrossRef]
  100. Saito, H.; Sakata, N.; Ishiga, T.; Ishiga, Y. Efficacy of RNA-spray-induced silencing of Phakopsora pachyrhizi chitin synthase genes to control soybean rust. J. Gen. Plant Pathol. 2022, 88, 203–206. [Google Scholar] [CrossRef]
  101. Cai, Q.; He, B.; Kogel, K.-H.; Jin, H. Cross-kingdom RNA trafficking and environmental RNAi—Nature’s blueprint for modern crop protection strategies. Curr. Opin. Microbiol. 2018, 46, 58–64. [Google Scholar] [CrossRef]
  102. 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. [Google Scholar] [CrossRef]
  103. Cheng, S.; Xie, X.; Xu, Y.; Zhang, C.; Wang, X.; Zhang, J.; Wang, Y. Genetic transformation of a fruit-specific, highly expressed stilbene synthase gene from chinese wild Vitis Quinquangularis. Planta 2016, 243, 1041–1053. [Google Scholar] [CrossRef]
  104. 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. [Google Scholar] [CrossRef]
  105. Both, M.; Eckert, S.E.; Csukai, M.; Müller, E.; Dimopoulos, G.; Spanu, P.D. Transcript profiles of Blumeria graminis development during infection reveal a cluster of genes that are potential virulence determinants. Mol. Plant Microbe Interact. 2005, 18, 125–133. [Google Scholar] [CrossRef]
  106. Pedersen, C.; van Themaat, E.V.L.; McGuffin, L.J.; Abbott, J.C.; Burgis, T.A.; Barton, G.; Bindschedler, L.V.; Lu, X.; Maekawa, T.; Wessling, R.; et al. Structure and evolution of barley powdery mildew effector candidates. BMC Genomics 2012, 13, 694. [Google Scholar] [CrossRef]
  107. Bourras, S.; Praz, C.R.; Spanu, P.D.; Keller, B. Cereal powdery mildew effectors: A complex toolbox for an obligate pathogen. Curr. Opin. Microbiol. 2018, 46, 26–33. [Google Scholar] [CrossRef] [PubMed]
  108. Fuller, K.K.; Rhodes, J.C. Protein kinase A and fungal virulence: A sinister side to a conserved nutrient sensing pathway. Virulence 2012, 3, 109–121. [Google Scholar] [CrossRef] [PubMed]
  109. Hu, S.; Zhou, X.; Gu, X.; Cao, S.; Wang, C.; Xu, J.-R. The cAMP-PKA pathway regulates growth, sexual and asexual differentiation, and pathogenesis in Fusarium graminearum. Mol. Plant Microbe Intract. 2014, 27, 557–566. [Google Scholar] [CrossRef] [PubMed]
  110. 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. [Google Scholar] [CrossRef]
  111. Wang, X.; Zhai, T.; Zhang, X.; Tang, C.; Zhuang, R.; Zhao, H.; Xu, Q.; Cheng, Y.; Wang, J.; Duplessis, S.; et al. Two stripe rust effectors impair wheat resistance by suppressing import of host Fe–S protein into chloroplasts. Plant Physiol. 2021, 187, 2530–2543. [Google Scholar] [CrossRef]
  112. 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. 2018, 16, 1013–1023. [Google Scholar] [CrossRef]
  113. Islam, M.T.; Sherif, S.M. RNAi-Based Biofungicides as a promising next-generation strategy for controlling devastating gray mold diseases. Int. J. Mol. Sci. 2020, 21, 2072. [Google Scholar] [CrossRef]
  114. Qiao, L.; Lan, C.; Capriotti, L.; Ah-Fong, A.; Nino Sanchez, J.; Hamby, R.; Heller, J.; Zhao, H.; Glass, N.L.; Judelson, H.S.; et al. Spray-induced gene silencing for disease control is dependent on the efficiency of pathogen RNA uptake. Plant Biotechnol. J. 2021, 19, 1756–1768. [Google Scholar] [CrossRef]
  115. Hu, D.; Chen, Z.; Zhang, C.; Ganiger, M. Reduction of Phakopsora pachyrhizi infection on soybean through host- and spray-induced gene silencing. Mol. Plant Pathol. 2020, 21, 794–807. [Google Scholar] [CrossRef]
  116. Degnan, R.M.; McTaggart, A.R.; Shuey, L.S.; Pame, L.J.S.; Smith, G.R.; Gardiner, D.M.; Nock, V.; Soffe, R.; Sale, S.; Garrill, A.; et al. Exogenous double-stranded RNA inhibits the infection physiology of rust fungi to reduce symptoms in planta. Mol. Plant Pathol. 2023, 24, 191–207. [Google Scholar] [CrossRef]
  117. Wang, M.; Jin, H. Spray-induced gene silencing: A powerful innovative strategy for crop protection. Trends Microbiol. 2017, 25, 4–6. [Google Scholar] [CrossRef]
  118. Taning, C.N.; Arpaia, S.; Christiaens, O.; Dietz-Pfeilstetter, A.; Jones, H.; Mezzetti, B.; Sabbadini, S.; Sorteberg, H.; Sweet, J.; Ventura, V.; et al. RNA-based biocontrol compounds: Current status and perspectives to reach the market. Pest Manag. Sci. 2020, 76, 841–845. [Google Scholar] [CrossRef] [PubMed]
  119. Landry, M.P.; Mitter, N. How nanocarriers delivering cargos in plants can change the GMO landscape. Nat. Nanotechnol. 2019, 14, 512–514. [Google Scholar] [CrossRef] [PubMed]
  120. Mitter, N.; Worrall, E.A.; Robinson, K.E.; Li, P.; Jain, R.G.; Taochy, C.; Fletcher, S.J.; Carroll, B.J.; Lu, G.Q.; Xu, Z.P. Clay nanosheets for topical delivery of RNAi for sustained protection against plant viruses. Nat. Plants 2017, 3, 16207. [Google Scholar] [CrossRef] [PubMed]
  121. Jain, R.G.; Fletcher, S.J.; Manzie, N.; Robinson, K.E.; Li, P.; Lu, E.; Brosnan, C.A.; Xu, Z.P.; Mitter, N. Foliar application of clay-delivered RNA interference for whitefly control. Nat. Plants 2022, 8, 535–548. [Google Scholar] [CrossRef] [PubMed]
  122. Mosa, M.A.; Youssef, K. Topical delivery of host induced RNAi silencing by layered double hydroxide nanosheets: An efficient tool to decipher pathogenicity gene function of Fusarium crown and root rot in tomato. Physiol. Mol. Plant Pathol. 2021, 115, 101684. [Google Scholar] [CrossRef]
  123. Niño-Sánchez, J.; Sambasivam, P.T.; Sawyer, A.; Hamby, R.; Chen, A.; Czislowski, E.; Li, P.; Manzie, N.; Gardiner, D.M.; Ford, R.; et al. BioClayTM prolongs RNA interference-mediated crop protection against Botrytis cinerea. J. Integr. Plant Biol. 2022, 64, 2187–2198. [Google Scholar] [CrossRef]
  124. Delgado-Martín, J.; Delgado-Olidén, A.; Velasco, L. Carbon dots boost dsRNA delivery in plants and increase local and systemic siRNA production. Int. J. Mol. Sci. 2022, 23, 5338. [Google Scholar] [CrossRef]
  125. Kostov, K.; Andonova-Lilova, B.; Smagghe, G. Inhibitory activity of carbon quantum dots against Phytophthora infestans and fungal plant pathogens and their effect on dsRNA-induced gene silencing. Biotechnol. Biotechnol. Equip. 2022, 36, 949–959. [Google Scholar] [CrossRef]
  126. Karny, A.; Zinger, A.; Kajal, A.; Shainsky-Roitman, J.; Schroeder, A. Therapeutic nanoparticles penetrate leaves and deliver nutrients to agricultural crops. Sci. Rep. 2018, 8, 7589. [Google Scholar] [CrossRef]
  127. Zhang, H.; Zhang, H.; Demirer, G.S.; González-Grandío, E.; Fan, C.; Landry, M.P. Engineering DNA nanostructures for siRNA delivery in plants. Nat. Protoc. 2020, 15, 3064–3087. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Interaction between a plant cell and fungal pathogen from the perspective of plant RNAi-mediated host-induced gene silencing. The arrows represent the flow of the gene silencing mechanism using RNAi. In the nucleus of a plant cell, dsRNA and hpRNA are produced as normal defense responses or from hairpin RNAs in transgenic RNAi plants (targeting a fungal gene). In addition, there are several biotechnology tools that allow the entry of exogenous dsRNAs or hpRNAs. These molecules of dsRNA and hpRNA can be processed by the DICER enzyme, creating siRNAs, which are integrated into the RNA-induced silencing complex (RISC), which contains an Argonaute protein (AGO), using them as templates for mRNA silencing. For the amplification of this silencing mechanism, there is an RNA-dependent RNA polymerase (RdRP) that can synthesize new dsRNAs using hybridized siRNA strands as primers. siRNAs produced in plant cells can be transported presumably by two types of transport (represented with question marks): via plant-derived extracellular vesicles (EVs) and plasma membrane-located transporters (PMTs). Inside the fungal cell, the mechanism of silencing works similarly to plant cells, producing the assembly of siRNAs with the RISC and inducing the silencing of specific fungal mRNAs.
Figure 1. Interaction between a plant cell and fungal pathogen from the perspective of plant RNAi-mediated host-induced gene silencing. The arrows represent the flow of the gene silencing mechanism using RNAi. In the nucleus of a plant cell, dsRNA and hpRNA are produced as normal defense responses or from hairpin RNAs in transgenic RNAi plants (targeting a fungal gene). In addition, there are several biotechnology tools that allow the entry of exogenous dsRNAs or hpRNAs. These molecules of dsRNA and hpRNA can be processed by the DICER enzyme, creating siRNAs, which are integrated into the RNA-induced silencing complex (RISC), which contains an Argonaute protein (AGO), using them as templates for mRNA silencing. For the amplification of this silencing mechanism, there is an RNA-dependent RNA polymerase (RdRP) that can synthesize new dsRNAs using hybridized siRNA strands as primers. siRNAs produced in plant cells can be transported presumably by two types of transport (represented with question marks): via plant-derived extracellular vesicles (EVs) and plasma membrane-located transporters (PMTs). Inside the fungal cell, the mechanism of silencing works similarly to plant cells, producing the assembly of siRNAs with the RISC and inducing the silencing of specific fungal mRNAs.
Ijms 24 09082 g001
Figure 2. (A) Powdery mildew symptoms observed on leaves and fruits of several crops. (1) Wheat (Triticum aestivum) leaf, (2) wine grape (Vitis vinifera) leaf, (3) soybean (Glycine max) leaves, (4) strawberry (Fragaria sp.), (5) tomato (Solanum lycopersicum) leaves, and (6) melon (Cucumis melo) leaves infected by Blumeria graminis, Erysiphe necator, Microsphaera diffusa, Podosphaera aphanis, Leveillula taurica, and Podosphaera xanthii, respectively. Pictures (1)–(6) were taken by Clemson University—USDA CES, Yuan-Min Shen (National Taiwan University), Daren Mueller (Homemade, Bugwood.org (accessed on 28 March 2023)), University of Hertfordshire, Scot Nelson (Homemade flickr.com (accessed on 28 March 2023), and by the authors of this review, respectively. (B) The typical powdery mildew life cycle is divided into two types of reproduction. Asexual reproduction is carried out by the release of conidium spores, which develop haustoria capable of acquiring nutrients from plant cells and giving rise to hyphae and conidiophores. Sexual reproduction occurs when two hyphae from opposing mating types form a chasmothecium capable of releasing an ascus with eight ascospores.
Figure 2. (A) Powdery mildew symptoms observed on leaves and fruits of several crops. (1) Wheat (Triticum aestivum) leaf, (2) wine grape (Vitis vinifera) leaf, (3) soybean (Glycine max) leaves, (4) strawberry (Fragaria sp.), (5) tomato (Solanum lycopersicum) leaves, and (6) melon (Cucumis melo) leaves infected by Blumeria graminis, Erysiphe necator, Microsphaera diffusa, Podosphaera aphanis, Leveillula taurica, and Podosphaera xanthii, respectively. Pictures (1)–(6) were taken by Clemson University—USDA CES, Yuan-Min Shen (National Taiwan University), Daren Mueller (Homemade, Bugwood.org (accessed on 28 March 2023)), University of Hertfordshire, Scot Nelson (Homemade flickr.com (accessed on 28 March 2023), and by the authors of this review, respectively. (B) The typical powdery mildew life cycle is divided into two types of reproduction. Asexual reproduction is carried out by the release of conidium spores, which develop haustoria capable of acquiring nutrients from plant cells and giving rise to hyphae and conidiophores. Sexual reproduction occurs when two hyphae from opposing mating types form a chasmothecium capable of releasing an ascus with eight ascospores.
Ijms 24 09082 g002
Figure 3. (A) Rust fungi symptoms observed on leaves of several crops: (1) coffee leaf (Coffea arabica), (2) barley leaf (Hordeum vulgare), (3) oat stem (Avena sativa), (4) black raspberry leaves (Rubus occidentalis), (5) glossy buckthorn (Frangula alnus), and (6) pear leaves (Pyrus spp.) infected by Hemileia vastatrix, Puccinia striiformis, P. graminis f. sp. avenae, Arthuriomyces peckianus, Puccinia coronata and Gymnosporangium sabinae, respectively. Pictures (1)–(6) were taken by Dr. Parthasarathy Seethapathy (Amrita School of Agricultural Sciences), Mary Burrows (Montana State University), Howard F. Schwartz (Colorado State University), Sandra Jensen (Cornell University), Milan Zubrik (Forest Research Institute—Slovakia; homemade, Bugwood.org (accessed on 28 March 2023)) and Sue Muller (Homemade MarylandBiodiversityProject.com (accessed on 28 March 2023)), respectively. (B) The typical rust fungal life cycle includes two types of hosts. First, aecial hosts are infected by haploid basidiospores ①, which generate pycnium as reproductive structures ②. Pycnium produce pycniospores with different polarities, which produce plasmogamy for generating haploid aeciospores ③. Second, aeciospores infect the telial host, in which urediniospores can be generated for asexual reproduction ④. In the telial host, teliospores are produced by karyogamy. Finally, basidiospores are produced by the meiosis of teliospores ⑤.
Figure 3. (A) Rust fungi symptoms observed on leaves of several crops: (1) coffee leaf (Coffea arabica), (2) barley leaf (Hordeum vulgare), (3) oat stem (Avena sativa), (4) black raspberry leaves (Rubus occidentalis), (5) glossy buckthorn (Frangula alnus), and (6) pear leaves (Pyrus spp.) infected by Hemileia vastatrix, Puccinia striiformis, P. graminis f. sp. avenae, Arthuriomyces peckianus, Puccinia coronata and Gymnosporangium sabinae, respectively. Pictures (1)–(6) were taken by Dr. Parthasarathy Seethapathy (Amrita School of Agricultural Sciences), Mary Burrows (Montana State University), Howard F. Schwartz (Colorado State University), Sandra Jensen (Cornell University), Milan Zubrik (Forest Research Institute—Slovakia; homemade, Bugwood.org (accessed on 28 March 2023)) and Sue Muller (Homemade MarylandBiodiversityProject.com (accessed on 28 March 2023)), respectively. (B) The typical rust fungal life cycle includes two types of hosts. First, aecial hosts are infected by haploid basidiospores ①, which generate pycnium as reproductive structures ②. Pycnium produce pycniospores with different polarities, which produce plasmogamy for generating haploid aeciospores ③. Second, aeciospores infect the telial host, in which urediniospores can be generated for asexual reproduction ④. In the telial host, teliospores are produced by karyogamy. Finally, basidiospores are produced by the meiosis of teliospores ⑤.
Ijms 24 09082 g003
Figure 4. Models showing RNA interference tools for plant and fungal pathogens. (A) VIGS: Virus RNAs (α, β, γ) are inoculated into plant cells. Inside the host, barley stripe mosaic virus (BSMV) is assembled. dsRNAs specific for the pathogen mRNA target are produced by the virus machinery. dsRNA processed by the silencing machinery silences the expression of the specific mRNA. (B) HIGS with micro-bombardment; hpRNAs join accelerated particles and are used for transitory transformation of plant cells. Inside the plant, hpRNAs are cut by the Dicer enzyme, and the sRNA derivate activates the RISC complex, which is able to hybridize with mRNA targets, allowing mRNA degradation by the ARGONAUTE enzyme. (C) ATM-HIGS: Agrobacterium tumefaciens transformed with Ti plasmid with a specific sequence of the pathogen target gene produces a transitory transformation of the plant. The Ti plasmid has sequences of vir genes. These vir genes encode several vir proteins responsible for transporting T-DNA into host cells. Inside the plant, T-DNA is introduced into the genome of the cells, allowing the production of hpRNAs. These hpRNAs are used to silence pathogen mRNAs. (D) dsRNA infiltration: dsRNA produced in vitro is introduced directly into the plant. These dsRNAs are specific for the silencing of pathogen mRNA targets. The dsRNA is cut into sRNAs; sRNAs joined to the RISC complex can hybridize with specific mRNAs being degraded by the ARGONAUTE enzyme.
Figure 4. Models showing RNA interference tools for plant and fungal pathogens. (A) VIGS: Virus RNAs (α, β, γ) are inoculated into plant cells. Inside the host, barley stripe mosaic virus (BSMV) is assembled. dsRNAs specific for the pathogen mRNA target are produced by the virus machinery. dsRNA processed by the silencing machinery silences the expression of the specific mRNA. (B) HIGS with micro-bombardment; hpRNAs join accelerated particles and are used for transitory transformation of plant cells. Inside the plant, hpRNAs are cut by the Dicer enzyme, and the sRNA derivate activates the RISC complex, which is able to hybridize with mRNA targets, allowing mRNA degradation by the ARGONAUTE enzyme. (C) ATM-HIGS: Agrobacterium tumefaciens transformed with Ti plasmid with a specific sequence of the pathogen target gene produces a transitory transformation of the plant. The Ti plasmid has sequences of vir genes. These vir genes encode several vir proteins responsible for transporting T-DNA into host cells. Inside the plant, T-DNA is introduced into the genome of the cells, allowing the production of hpRNAs. These hpRNAs are used to silence pathogen mRNAs. (D) dsRNA infiltration: dsRNA produced in vitro is introduced directly into the plant. These dsRNAs are specific for the silencing of pathogen mRNA targets. The dsRNA is cut into sRNAs; sRNAs joined to the RISC complex can hybridize with specific mRNAs being degraded by the ARGONAUTE enzyme.
Ijms 24 09082 g004
Figure 5. Molecular process scheme of SIGS assays for fungal control. The designed dsRNAs are sprayed onto the host plant leaf. The first way of uptake is by the plant ① with the subsequent activation of the DICER-RISC-mediated silencing system ②–③ degrading mRNA targets of the pathogen ④; however, some fungal structures, such as haustorium, can presumably uptake small molecules as siRNAs and activate the DICER-RISC complex inside the fungal cell as it is indicated in question marks.
Figure 5. Molecular process scheme of SIGS assays for fungal control. The designed dsRNAs are sprayed onto the host plant leaf. The first way of uptake is by the plant ① with the subsequent activation of the DICER-RISC-mediated silencing system ②–③ degrading mRNA targets of the pathogen ④; however, some fungal structures, such as haustorium, can presumably uptake small molecules as siRNAs and activate the DICER-RISC complex inside the fungal cell as it is indicated in question marks.
Ijms 24 09082 g005
Table 1. Virus-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Table 1. Virus-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Plant HostPathogenTarget GenePossible Gene FunctionApplicationPhenotypeReferences
Hordeum vulgareBlumeria graminis f. sp. hordeiGTF1Cell wall elongation and virulence factorVirus inoculation by rubbing of barley first leavesReduction in haustorium formation[53]
GTF2
Triticum aestivumPuccinia striiformis f. sp. triticiPSTha12J12Predicted secreted proteinVirus inoculation by rubbing wheat leavesReduction in the expression patternsof the fungal genes[62]
PSTha5A23
PSTha12H2
PSTha2A5
PSTha9F18
PSTha5A1Predicted to code for a chitinase protein
PSTha12O3Homologous to Uromycesfabae hexose trans-porters
PsCNA1Calcineurin A-like protein (CNA1)Slower elongation of fungal hyphae and reduction of the production of uredospore[63]
PsCNB1Calcineurin B-like protein (CNB1)
PsSRPKLProtein kinaseReduction of fungal growth and increases of ROS accumulation in host cells[64]
PsANTAdenine nucleotide translocaseAttenuated the growth and development of virulent Pst at the early infection stage[65]
PsSOD1Zn-only superoxide dismutaseReduction of the virulence-associated with ROS accumulation[66]
PsRanSmall GTP-binding proteinReduction of the number of haustoria and the length of infection hyphae[67]
PsFUZ7MAPK kinaseReduction of initial haustorium formation and elongation of secondary hyphae[68]
PstSTE12Transcription factorReduction in the growth and spread of hyphae in Pst and weakened the virulence of Pst on wheat[69]
T. aestivumP. striiformis f. sp. triticiPsCPK1PKA catalytic subunitVirus inoculation by rubbing wheat leavesReduction in the length of infection hyphae and disease phenotype[70]
PstMCM1-1MADX-box transcription factorReduction of hyphal extension and haustorium formation[71]
PsKPP4MAPK kinaseReduction of haustorium number[72]
Pst_8713Suppresses host defenses and contributes to the pathogenicity of PstReduction of haustorium number[73]
PstGSRE1Effector to defeat ROS-associated plant defense by modulating the subcellular compartment of a host immune regulatorReduction in sporulation and in the fungi biomass[74]
Puccinia triticinaPtCYC1CyclophilinReduction in fungal growth and disease symptoms[75]
PtMAPK1MAP kinase
PtCNBCalcineurin regulatory subunit
Puccinia graminis f. sp. triticiPgt-IaaMTryptophan mono-oxygenaseReduction in fungal growth and in the size of uredinia[76]
PGTG_01136Predicted glycolytic enzymeReduction in fungal growth and in the size of uredinia[77]
PGTG_01215Probably involved in cellular carbohydrate or sugar metabolism
PGTG_03478
PGTG_14350Hypothetical secreted protein with homology to periplasmic components of prokaryotic transport systems
PGTG_10731Hypothetical proteins
PGTG_12890
PGTG_01304Protein involved in thiazole biosynthesis
PGTG_16914Amino acid permease
PGTG_03590Secreted protein
Pgt-IaaMTryptophan 2-monooxygenase enzyme
Table 2. Host-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Table 2. Host-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Plant HostPathogenTarget GenePossible Gene FunctionApplicationPhenotypeReferences
H. vulgareBlumeria graminis f. sp. hordeiAvra10Virulence effectorMicroprojectile bombardmentReduction in haustorium formation[53]
BEC1054Ribonuclease-like proteinReduction in haustorium formation[54]
BEC1011
BEC1019Metalloprotease
BEC1005Endo β1-3 glucanase
CSEP0055Effector involved in secondary penetration eventsReduction in haustorium formation[80]
CSEP0105Effector proteinsReduction in haustorium formation[81]
CSEP0162
CSEP0027Interacts with barley HvCAT1 to regulate the host immunity to promote fungal virulenceReduction in haustoria formation[82]
CSEP0007Possibly involved in penetration and/or establishment of primary haustoriaReduction in haustoria formation[83]
CSEP0025
CSEP0128
CSEP0247
CSEP0345
CSEP0420
CSEP0422
CSEP0081Candidate Secreted Effector ProteinsMicroprojectile bombardmentReduction in fungal growth and in haustorium formation[84]
CSEP0254
CSEP0139Suppressed cell death triggered by BAX and NtMEK2DDReduction in haustoria formation[85]
CSEP0182
Table 3. Agrobacterium tumefaciens-mediated host-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Table 3. Agrobacterium tumefaciens-mediated host-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Plant HostPathogenTarget GenePossible Gene FunctionApplicationPhenotypeReferences
T. aestivumPuccinia triticina
Puccinia graminis and
Puccinia striiformis
PtCYC1CyclophilinAgroinfiltration through the abaxial surface of wheat seedling leavesReduction in fungal growth and sporulation[75]
PtMAPK1MAP kinase
PtCNBCalcineurin regulatory subunit
Cucumis meloPodosphaera xanthiiPEC007Candidate effectorAgroinfiltration of melon cotyledonsReduction of fungal growth and increasing of the production of hydrogen peroxide by host cells[55]
PEC009
PEC034
PEC032α-Mannosidase
PEC019Phospholipid-binding protein
PEC054Cellulose-binding protein
PEC1666Chitinase activityReduction of fungal growth and increasing of the production of hydrogen peroxide by host cells[92]
PEC1961
PEC2158
PEC5191
PHEC27213Lytic polysaccharide mono-oxygenase (LPMO) prevents the activation of chitin-triggered immunityReduction of fungal growth and increasing production of hydrogen peroxide by host cells[93]
PxCDAchitin deacetylaseReduction of fungal growth and increasing production of hydrogen peroxide by host cells[94]
Table 4. dsRNA-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Table 4. dsRNA-induced gene silencing method used for gene function analyses in rust and powdery mildew fungi.
Plant HostPathogenTarget GenePossible Gene FunctionApplicationPhenotypeReferences
Pisum sativumErysiphe pisiEpCSEP001Virulence factorsSecond leaves of pea plants were infiltrated with 100 parts per million (ppm) EpCSEP/CSP-dsRNAReduction in disease symptoms[99]
EpCSEP009
EpCSP083
C. meloPodosphaera xanthiiPxCNAP1048Presumably involved in glycosylationMelon cotyledons were infiltrated with dsRNA solutions of the different target genes in concentrations between 100 and 1000 ng ml−1Reduction in fungal growth and disease symptoms[29]
PxCNAP10905Presumably involved in respiration
PxCNAP30520
PxCNAP8878
PxCNAP9066
PxCNAP948Presumably involved in efflux transport
PxTUB2Involved in β-tubulin synthesis
PxCYP51Involved in ergosterol synthesis
Glycine maxPhakopsora pachyrhiziCHSInvolved in chitin synthasesSoybean plants were infiltrated with 10 ng ml−1 of dsCHSReduction in fungal growth and in the number of urediniospores[100]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Padilla-Roji, I.; Ruiz-Jiménez, L.; Bakhat, N.; Vielba-Fernández, A.; Pérez-García, A.; Fernández-Ortuño, D. RNAi Technology: A New Path for the Research and Management of Obligate Biotrophic Phytopathogenic Fungi. Int. J. Mol. Sci. 2023, 24, 9082. https://doi.org/10.3390/ijms24109082

AMA Style

Padilla-Roji I, Ruiz-Jiménez L, Bakhat N, Vielba-Fernández A, Pérez-García A, Fernández-Ortuño D. RNAi Technology: A New Path for the Research and Management of Obligate Biotrophic Phytopathogenic Fungi. International Journal of Molecular Sciences. 2023; 24(10):9082. https://doi.org/10.3390/ijms24109082

Chicago/Turabian Style

Padilla-Roji, Isabel, Laura Ruiz-Jiménez, Nisrine Bakhat, Alejandra Vielba-Fernández, Alejandro Pérez-García, and Dolores Fernández-Ortuño. 2023. "RNAi Technology: A New Path for the Research and Management of Obligate Biotrophic Phytopathogenic Fungi" International Journal of Molecular Sciences 24, no. 10: 9082. https://doi.org/10.3390/ijms24109082

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop