Next Article in Journal
Evaluation of the Productivity and Potential Utilization of Artemisia dubia Plant Biomass for Energy Conversion
Previous Article in Journal
In-Depth Characterization of bZIP Genes in the Context of Endoplasmic Reticulum (ER) Stress in Brassica campestris ssp. chinensis
Previous Article in Special Issue
Comparing Fungal Sensitivity to Isothiocyanate Products on Different Botrytis spp.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Update on the Basic Understanding of Fusarium graminearum Virulence Factors in Common Wheat Research

by
Zeeshan Ali Buttar
1,2,
Mengquan Cheng
1,2,
Panqin Wei
1,2,
Ziwei Zhang
1,2,
Chunlei Lv
1,2,
Chenjia Zhu
1,2,
Nida Fatima Ali
3,
Guozhang Kang
1,
Daowen Wang
1,2,* and
Kunpu Zhang
1,2,*
1
State Key Laboratory of Wheat and Maize Crop Science, Center for Crop Genome Engineering, College of Agronomy, Henan Agricultural University, Zhengzhou 450002, China
2
The Shennong Laboratory, Zhengzhou 450002, China
3
Department of Plant Biotechnology, Atta-Ur-Rehman School of Applied Biosciences (ASAB), National University of Science and Technology, Islamabad 44000, Pakistan
*
Authors to whom correspondence should be addressed.
Plants 2024, 13(8), 1159; https://doi.org/10.3390/plants13081159
Submission received: 28 December 2023 / Revised: 16 April 2024 / Accepted: 19 April 2024 / Published: 22 April 2024
(This article belongs to the Special Issue Fungus and Plant Interactions: Volume II)

Abstract

:
Wheat is one of the most important food crops, both in China and worldwide. Wheat production is facing extreme stresses posed by different diseases, including Fusarium head blight (FHB), which has recently become an increasingly serious concerns. FHB is one of the most significant and destructive diseases affecting wheat crops all over the world. Recent advancements in genomic tools provide a new avenue for the study of virulence factors in relation to the host plants. The current review focuses on recent progress in the study of different strains of Fusarium infection. The presence of genome-wide repeat-induced point (RIP) mutations causes genomic mutations, eventually leading to host plant susceptibility against Fusarium invasion. Furthermore, effector proteins disrupt the host plant resistance mechanism. In this study, we proposed systematic modification of the host genome using modern biological tools to facilitate plant resistance against foreign invasion. We also suggested a number of scientific strategies, such as gene cloning, developing more powerful functional markers, and using haplotype marker-assisted selection, to further improve FHB resistance and associated breeding methods.

1. Introduction

Bleached ears caused by the fungus Fusarium graminearum have emerged in luscious green fields of wheat, threatening wheat producers around the world [1]. After infection, the fungus spreads rapidly in the ear; continues to multiply and grow during the grain filling and maturity of wheat; and produces a variety of toxins in wheat grains, including deoxynivalenol (DON), nivalenol (NIV), and zearalenol (ZEN). Once these toxins enter the human body or livestock, they cause the body’s immunity to decline, leading to teratogenesis, cancer, abortion in pregnant women, and other pathogenic effects, thus posing serious harm to the health of humans and livestock. Historically, the first outbreak of this fungus was reported in 1884 in England, followed by America, Europe, Australia, and South Africa [2,3,4]. An estimated loss of USD 3 billion throughout the 1990s was documented in North America. Since then, approximately 28 million metric tons of wheat grain has been infected, accounting for almost USD 5.6 billion [1,5]. From 2004 to 2012, the annual economic losses ranged from CAD 1 million to 9 million in Alberta, Canada (Government of Alberta, 2015).
Immature cereal grains become infected with the fungus at the anthesis stage. Thereafter, fungal hyphae spread within the spikelet through the vascular bundles of the rachis [6], thus promoting infection. In this case, one of the main factors influencing a plant’s susceptibility is thought to be the pathogen’s capacity to subvert its host’s physiological functions, including defense mechanisms, physiology, and basic metabolism, in order to make use of its resources. The two partners engage in intricate molecular cross-talk during these interactions, which includes the release of effectors, small secreted proteins with the ability to change the structure of host cells and target particular functions in host tissues which are also known as susceptibility factors. In response, the host plant activates a two-layered immune response [7]. The first layer is called pathogen-triggered immunity (PTI), which is activated when the plant recognizes the pathogen/microbial-associated molecular pattern (PAMP/MAMP) on the surface of the pathogen through the use of its pattern recognition receptors (PRRs) [8,9,10]. In order to curb this primary response, the pathogen release an arsenal of molecules, collectively called effector proteins. These molecules work as potential agents within the host to augment the capabilities through interfering with different cellular processes, such as signaling, cellular adherence, transcription, vesicular trafficking, membrane biogenesis, apoptosis, and metabolism [11]. However, some of these effectors are directly or indirectly detected by plants and deploy R-genes encoded with R-proteins, leading to rapid cell death; this is termed as the hypersensitive response (HR) or effector-triggered immunity (ETI) [11,12]. ETI is similar to the process of apoptosis in animals [13]. In contrast, resistance develops as depicted in Flor’s gene-for-gene model through the presence of similar R and effector proteins [14,15]. However, the proper delivery of these effector molecules is critical for the success of a pathogen. A small number of fungi, such as Fusarium, Blumeria graminis, Puccina recondite, Puccinia graminis, Puccina striformis, Septoria nodorum, and Septoria tritici, have been studied at this level. However, it is still difficult to uncover the multiple strategies of the mode of action of effectors, such as how the fungi maintain their effector protein evolutionary mechanism and what specific mechanism supports the pathogen in sustaining this evolutionary machinery in contrast with host plants.
It has been suggested that repeat-induced point mutation (RIP) has an impact on genome duplication through gene duplication. Therefore, RIP is considered the most critical factor in the evolution of new functions [16,17]. In addition, epigenetic changes through genome-wide RIP mechanisms are a major weapon through which pathogens evolve in response to host plants. However, there is still a long way to go to understand the mechanism behind RIP. In this study, we specifically discuss and update the understanding of Fusarium graminearum virulence factors in common wheat, which may lead to overcoming the unknown strategies with which fungi maintain their evolutionary mechanism for completion with their host plant.

2. Distinguished Features of Fusarium graminearum Life Cycle

Most of the F. graminearum life cycle is linked with the host plant (see Figure 1), and the fungus is haploid throughout much of its life cycle. The production of hyphae represents a bottleneck for sexual development within binucleate cells. Further, F. graminearum belongs to the phylum Ascomycota, which represents the binucleate phase known as the dikaryotic phase, in which two genetically separate nuclei remain coupled as new cells grow. It also possesses a homothallic nature, which means that it may produce sexual spores (ascospores) without the assistance of a sexually unique partner. This leads to the formation of binucleate cells (or “two nuclei cells”), which are genetically identical. Homothallism is caused by the haploid genome, including Mat1-1 and Mat1-2 genes from both mating types [17,18,19,20]. For example, F. graminearum colonizes the surfaces of wheat plants through the use of specialized unbranched hyphae known as runner hyphae (RH). Therefore, multi-celled complex appressoria known as infection cushions (IC) are developed, which accumulate an arsenal of (proven and putative) virulence factors to facilitate the invasion of epidermal cells [21,22].

3. Fusarium graminearum Germination Leading Up to Infection

The fungal plant pathogen initiates its first interaction with the host through spore germination. Spore germlings travel across the plant surface and utilize different invasion strategies to penetrate the host plant surfaces. These penetration strategies comprise pressurized melanized appressoria, which assists in physically punching through the plant cuticle, and non-melanized appressoria, which penetrates with the help of enzymes or cuticle damage to breach the plant surface. Further, the reactive oxygen species (ROS) level plays a significant role in the pathogen–host interaction. Thus, the KatG2-mRFP genes deliberately control the ROS level on the cell wall of invading hyphal cells. The pathogen delineates a change in infection by shifting the ROS levels through temporal and spatial regulation of KatG2 to counteract the oxidative stress generated by host plant cells [20]. In contrast, over-expression of DOHH promotes ROS and sexual reproduction [23]. FgMet3 and FgMet14, which are localized to the cytoplasm, have been identified in the synthesis of methionine and cysteine, and are further recognized to play a functional role in penetrability within the host plant [24]. Transcriptomic analysis of the Tri genes Tri5, Tri6, Tri8, Tri9, and Tri14 indicated a significant lack of expression during conidial germination and plant penetration, whereas Tri1, Tri3, Tri4, Tri7B, Tr10, Tri11, and Tri12 exhibited expression only at the first polar growth stage in wheat and barley. As such, it was deduced that the biosynthesis of DON by F. graminearum is not fully functional until after penetration [25]. The accumulation of DON in wheat seeds poses threats to human and animal health. In this case, linoleic acid (LA; cis-9, cis-12 C18:2) has been shown to have a significant role in wheat’s resistance to F. graminearum infection. It has been suggested that LA reinforces the cuticle, which acts as a barrier to pathogen entry [26]. Disruption of FgLAI12 during wheat spike infection increases LA and SA levels to a higher extent in resistant wheat lines than in susceptible lines [27,28]. Therefore, FgLAI12 is critical for mycelial growth and virulence in wheat. In addition, the fg3_54 gene has been shown to facilitate the ability to infect wheat spikes through cell-to-cell penetration [29]. It has been deduced that effector proteins play significant roles in promoting pathogen manipulation in the host plant.

4. Role of Effectors in Colonization of Host Plant

Effector proteins allow the pathogen to colonize the host plant through various mechanisms. These genes basically prevent recognition by the host, interfere with phytohormone defense pathways, regulate host gene expression, and influence host protein trafficking [30,31]; see Table S1. Effectors initially suppress the PAMPs response; for example, ROS production is the preliminary “PTI” signal against pathogen invasion, and a large number of ROS-related genes have been found to be activated in F. graminearum at 16 h after host invasion [32]. A detailed model of the ROS genes expressed at 16 hai [33] is shown in Figure 2. Similarly, peroxisomes are organelles that have been identified to play a role in ROS detoxification. F. graminearum secretes an effector, Osp24, which induces degradation of the wheat TaSnRK1α kinase to promote disease, while an orphan wheat protein, TaFROG1, can compete with Osp24 for binding to TaSnRK1α and protect it from degradation [34]. The PEX5 and PEX6 genes from F. graminearum play a role in ROS detoxification, and deletion of PEX5 and PEX6 results in accumulation of ROS followed by necrotic cell death [35]. It has been suggested that peroxisomes are key organelles that balance ROS levels in filamentous fungi. Bioinformatics studies have identified 600 effectors secreted by F. graminearum [36], where 30 effector proteins showed interaction with a small cysteine-rich protein (SCPP) which contained N-terminal signal peptides and lacked transmembrane domains [36]. PRRs first recognize effectors to produce PTI, the failure of which results in a shift to the second layer of defense, ETI. F. graminearum has been identified to secrete plant-cell-wall-degrading enzymes composed of pectate lyases (cleave pectin), such as “FG02386, FG03131, FG03483, FG03908, and FG04864”, and xylanases (degrade xylan) such as “FG00184, FG07639, and FG11304”, which have significant contributions in the hemicellulose of monocot cell walls. These enzymes have been predicted to be involved as effector molecules that trigger host plant defense responses [37]. Therefore, they have important roles regarding acquisition and penetration of plant tissues [38]. F. graminearum has been shown to possess a higher number of plant necrosis-inducing proteins (or NIPs, also called NLPs), in contrast to N. crassa and A. nidulans. The first of the NLPs was a 24 KDa protein isolated from F. oxysporum, which has the capability to induce necrosis and ethylene biosynthesis in numerous dicotyledonous plants [39].

5. Host Plant Resistance Mechanism against Effector Invasion

Host plants have evolved a sophisticated two-layered defense system. PAMPs are recognized by PRRs, triggering the first PTI response, followed by a second layer comprising the cytoplasmic ETI response; failure of these responses results in ETS [41]. The plant susceptibility “S” factor has gained significant importance. The removal of “S” could lead toward resistance, as it indirectly supports the pathogen’s virulence mechanism [42,43,44,45]. Pathogens bombard a range of effector proteins into the host tissue through the formation of haustoria within the host plant, which are surrounded by a plant-derived extra-haustorial membrane (EHM) [42,46]. It has been suggested that the extra-haustorial membrane conducts effector screening before their release into the host plants. For example, in F. graminearum, 22 hexose transporters with various roles and specificities have been predicted, and have yet to be investigated [7,33,39,46]. In response, the host plant initiates NADPH-based ROS production, resulting in downstream signaling of PRRs and inducing Ca2+ influx [47]. Therefore, defense genes are named resistance (R) genes, which are deployed to protect against one or more pathogenic strains. R-genes most commonly encode NLRs, and resistance occurs when the NLR proteins identify an effector and subsequently stimulate efficient ETI. NLR genes have been observed, to some extent, in all resistance mechanisms. However, it is still a matter of debate whether all NLR genes in plant species can be genetically exposed as R genes. A typical NLR has a nucleotide-binding, Apaf1-resistant, CED4 (NB-ARC) central domain, as well as a leucine-rich repeat (LRR) C-terminal domain. Based on the N-terminal domain, NLRs can be further classified into three major classes: CC-type NLRs (RNLs), coiled-coil domain CC-NLRs (CNLs), and interleukin-1 receptor domain TIR-NLRs (TNLs) (112). In the case that NLRs bind with an effector or sense a modification of an effector target, a conformational change opens the P-loop of the NB-ARC domain to force the conversion of ADP to ATP. It has been speculated that this might be the mechanism used by NLR to trigger ETI [48]. The effectors that are recognized by R genes or NLRs are known as AVR genes or proteins. For example, F. oxysporum-secreted Avr2 was recognized by the tomato 1-2 (NB-LRR)-type resistance protein. However, point mutations prevent recognition [49]. Similarly, the introduction of a 1-3-mediated resistance gene in tomato yielded resistance against F. oxysporum (AVR1) [50]. In addition, it is likely that Osp24 suppression provides protection against FHB through interaction with TaNACL-D1. Two orphan proteins, TaFROG and Osp24, are found in hosts and pathogens, respectively. Therefore, during fungal–plant interactions, each might experience co-evolution. To the best of our knowledge, no other pathosystems have been reported regarding the active adoption of competing orphan proteins in either host plants or fungal pathogens. As they encode orphan proteins that are specifically expressed during infection, and expression-engineered TaFROG alleles with stronger interactions with TaSnRK1α or silencing OSP24 may improve resistance against F. graminearum without yield penalties [34]. There remains significant pending work to determine the F. graminearum Avr-genes.

5.1. Effectors Inhibit R-Genes Expression through Chromatin Remodeling

Morphological changes are a very common and effective strategy used by pathogens to survive in the host plant. Therefore, during interaction with their host, pathogenic fungi undergo an array of morphological changes that are closely associated with numerous patterns in order to regulate their gene expression and biological processes [51]. In eukaryotic organisms, the genetic information comprises a specific nucleoprotein complex that is packed with an array of nucleosomes, known as chromatin. This is wrapped around a core of four histone proteins, namely, H2A, H2B, H3, and H4 [51,52]. In order to sense, environmental stimuli are integrated through epigenetic processes including chromatin remodeling, either allowing or inhibiting gene expression at the molecular level [52]. F. graminearum is a well-known hemi-biotroph fungus, and is the causative agent of FHB in cereal crops. For successful pathogen infection, each stage comprises specific genetic modifications [53], which are strictly associated with certain virulence mechanisms. These modifications are responsible for different post-translational modifications (PTMs) of histones, including acetylation, deacetylation, methylation, phosphorylation, ubiquitination, etc. Coordination between various histone modifiers defines a specific chromatin state of the genome [54]. Similarly, the immune response against pathogen infection requires extensive transcriptional reprogramming. Histone acetylation is considered to play a vital role in transcriptional regulation; for example, the cytoplasmic effector PsAvh23 of the pathogen Phytophthora sojae has been identified as a modulator of histone acetyltransferase (HAT) in plants [55], resulting in host susceptibility through suppression of the immune response. Hence, histone modification (except in the case of methylation) could lead to alterations in the histone–histone or histone–DNA interactions. Furthermore, similar effects could be observed with respect to other histone proteins that could influence various changes in the function and structure of chromatin.
Previous successful work has identified several signaling pathways and chromatin modifiers as being essential for fungal pathogenicity. Histone acetyltransferases (HATs), “addition of acetyl group”, and histone deacetylases (HDACs), “removal of acetyl group”, are the well-known PTMs that regulate the pattern of gene expression, and are schematically represented in Figure 3A,B. The balancing act between these responses is important for appropriate cellular development and function [55]. HATs can be sub-divided into five sub-families based on conserved motifs: MYST (MOZ, Tip60, Sas2, Ybf2/Sas3), GNAT (Gcn5-related N-acetyltransferases), p300/CBP, basal transcription factor (including TFIID), and nuclear receptor cofactors. They can be further divided based on their cellular location in nuclear type-A or cytoplasmic type-B HATs, regardless of which family they belong to. A summary of their sub-cellular locations with Go-annotation is provided in Table 1.
In contrast, HDACs reverse the activity of HATs through removing acetylation on the ε-amino group of lysine residues. Thus, HDACs have been observed to be transcriptional repressors. They restore the positive charge of lysine, rendering the underlying DNA sequences relatively inaccessible to transcriptional machinery through stabilization of the local chromatin structure, as explained in Figure 3A. HDACs have been classified into four sub-classes based on phylogenetic analysis and sequence homology [56]. In previous work, successful degradation of F. graminearum histone acetylation at H2BK11, H3K11, H3K14, H3K18, and H3K27 has significantly reduced fungal growth, virulence, and mycotoxin biosynthesis [51,57]. Considering the results of previous studies, the transcription factor FgPacC plays a significant role in protecting the fungus from iron toxicity through direct binding of promoters and inhibition of SAGA activity [58]; however, there is still a need to carry out significant research to observe the accurate mechanisms through which the morphological changes and virulence factors of the pathogen are controlled. Furthermore, several defense factors, including phenolics, alkaloids, polyacetylenes, hydrogen peroxide (H2O2), and a series of pathogenesis-related (PR) compounds, may be produced by the plant. Therefore, epigenetic modification within the host genome through pathogen invasion may significantly suppress the production of plant defense proteins. For example, in the case of Blumeria graminis f. sp tritici (Bgt) causing powdery mildew in bread wheat (Triticum aestivum L.), significant improvements in TaPR1, TaPR2, TaPR5, and TaWRKY45 expression were observed through silencing of TaHBT701, TaHDA6, and TaHOS15 [59]. It can be speculated that their resistance mechanism might be suppressed by F. graminearum. It has been observed that MsDef1 and MtDef4 are produced in Medicago spp. to resist against F. graminearum growth [60].

5.2. Effector Protein Evolution through Repeat-Induced Point Mutation

Transposable elements (TEs) are considered a major factor in genome expansion, and are spread through the genome through a self-copying mechanism [61,62]. At the same time, TEs pose a severe threat to the pathogen’s genome, as the effector genes are commonly located in TE-rich regions, with epigenetic regulations observed in an extensive range of crop pathogens [62]. Therefore, the expression of effector genes is mostly controlled through variation in the heterochromatin state upon host–pathogen infection. In this scenario, to mitigate the adverse effects of “self-repeat” sequencing, numerous filamentous fungi species use a genome-wide defense system known as the repeat-induced point (RIP) mutation [63]. Further, the point mutation stops the translation of DNA such as the amino acid glycine, which may be changed to a stop codon, causing the proteins to be unable to complete the intended tasks. The problem arises during the processes of transcription and replication of DNA. These changes stop the cell from reproducing and thus lead to the death of the cell. For example, the Leptoshaeria maculans genome comprises one-third AT and contains effector genes and TEs families, both of which are affected by RIP. This novel mechanism for effector genes promotes rapid sequence diversification and allows the fungus to adapt rapidly to novel host-derived constraints [64]. Therefore, it is considered to be one of the main fungal-specific genome defense mechanisms. Point mutation of G: C to A: T at a very high rate is introduced, leading to non-functional TE copies [19,65]. By consequently reducing the GC content in the affected sequence, further RIP promotes a large block of evolution through gene duplication [66]. However, it restricts pathogen effector evolution, which is mostly prone to duplication due to the repetitive nature of their chromosomal locations [67]. Selker et al. conducted a pioneering study in Neurospora crassa to identify RIP [68,69]. Some key characteristics of RIP have been defined in N. crassa [70,71], as demonstrated in Figure 4. Hokyoung et al. confirmed RIP in Gibberella zea (anamorph: F. graminearum) [72]. F. graminearum has a homothallic (self-fertile) nature, which differentiates it from other filamentous fungi, and it rarely out-crosses to other strains, thus limiting it to gaining new repeats [73]. In addition, the presence of a genome-wide RIP mechanism limits F. graminearum to acquiring repetitive sequences. RIP permanently mutates the cytosine in duplicated motifs, inducing adenine- and thymine-biased transition mutation in the target sequences [74]. Likewise, transposon mutation during the sexual cycle of F. graminearum introduces C: G to T: A [75,76]. Further, RIP has been explored through epigenetic silencing in N. crassa. The entire process occurs during sexual reproduction, prior to karyogamy and meiosis, when two copies of the genome are still present in the dikaryotic cell. Therefore, RIP is thought to be the most clear-cut example of a genome defense mechanism, as it has no other known purpose [77]. The genomes of a number of species or subspecies of more recent and ancient outgroups that shared a common ancestor will be sequenced in order to test the validity of the proposed evolution scenario of RIP and provide additional information on the origin of the effectors, genome invasion by TEs, and the subsequent effect on effector generation and diversification.

6. Important Unresolved Questions

The host’s innate immune system mostly recognizes cysteine-rich proteins. In the case of F. graminearum, 76 peptides with less than 200 aa contain at least 4% cysteine residues, which may be speculated as the possible reason for PTI failure against F. graminearum. As previously identified, avirulence effectors of fungi are small, cysteine-rich proteins [79,80]. Therefore, further investigation is required in order to address related questions. In previous work, wheat histone (H1-H4) has been found to be involved in the F. graminearum resistance mechanism [59,81,82], which raises a question related to the histone–histone protein interactions taking place between F. graminearum and wheat histone protein, as well as how this may stimulate wheat susceptibility, which remains a very interesting and elusive question.

6.1. Molecular Pathogenicity

Necrotrophic pathogens feed and live on dying tissue, in contrast to biotrophic pathogens. Their resistance mechanism has been investigated in association with the wall-associated kinase class of receptors, which drives the resistance pathway against the biotrophic pathogen. Necrotrophic pathogens such as Phaeosphaeria nodorum have been observed to hijack the host molecular pathways involved in the resistance mechanism, demonstrating the complex nature of susceptibility and resistance in the interactions between necrotrophic or biotrophic pathogens and plants [83]. Therefore, to gain access to the host cell, pathogens either use natural opening (e.g., stomata), wounds, the aid of vectors, or utilize an arsenal of cell-wall-degrading enzymes (CWDEs) to deconstruct the structural components of host CW, assisting penetration and diffusion in the host tissues, and at the same time providing carbon sources and promoting leakage of nutrients from the protosplast [83]. Therefore, we can speculate the possible involvement of F. graminearum in hijacking the host wall-associated kinase class through the release of cell-wall-degrading enzymes [84,85,86], resulting in a lack of immune response against F. graminearum infection. Further, co-variation of wheat and fungal protein accumulation at different time points indicated that F. graminearum effector proteins intercept the molecular mechanisms, thus determining the fate of the interaction [85]. WRKY transcription factors are members of the large protein family. In a previous study, TaWRKY45 was found to be up-regulated in response to benzothiadiazole (BTH), a plant immune system strengthener [86]. Here, we propose that histone modification from HAT to HDAC suppressed TaWRKY45 expression, resulting in a lack of immunity response against FHB. The histone–histone interaction, as seen in Figure 3B, caused modifications in wheat histone (H1-H4) proteins, resulting in susceptibility of the wheat. Furthermore, histone proteins have been linked with a resistance mechanism against F. graminearum [86].

6.2. How Can We Develop Resistant Cultivars More Efficiently?

Previous studies have speculated on the negative correlation between FHB severity, plant height, and anther exclusion. It has been suggested that tall genotypes without anthers could have some level of resistance against FHB [87,88,89,90,91]. Recent studies have uncovered the underlying mechanism of glutathione S-transferase (GST) encoded by Fhb7. The Fhb7 is conserved in Epichloë species and can detoxify the trichothecene mycotoxins secreted by Fusarium species. Thus, integration of Fhb7 into a plant genome could be beneficial in eliminating the need for symbiotic association per se. The finding of Fhb7-mediated resistance to both FHB and crown rot diseases further emphasizes the importance of the use of Thinopyrum elongatum, a species used in distant hybridization breeding of wheat. Therefore, further thorough investigation is required when using genetic material in resistance breeding, as shown in Figure 5.

6.3. Climate Change on FHB Disease and Solution

In addition to the susceptible host, multiple environmental factors, including location, temperature, humidity, salinity, flooding, and drought, play vital roles in regulating the pathogenicity of FHB. In addition to this, the varying atmospheric CO2 levels also affect the spread of FHB. The F. graminearum infection is promoted by environmental changes that are related to the induction of DON production. A high level of DON production related to sucrose, 1-kestose, and nystose was observed in F. graminearum. Further, the carbon catabolic repression negatively regulates the expression of genes required for utilizing carbon sources. Therefore, predicted binding sites of FgCreA are present in promoters of TRI1, TRI3-TRI8, TRI10, TRI12, and TRI101 [81,86]. However, the extent of the effect of FgCreA on the regulation of DON biosynthesis is still unknown in relation to carbon and other environmental factors.

6.4. Strengthen the Study of Molecular Mechanisms of Pathogenic Bacteria’s Pathogenesis

Strong evidence of susceptible sites and genes may further strengthen the investigation of the molecular mechanism underlying the pathogenesis, growth, and development of pathogenic bacteria. Therefore, effective use can be made of gene editing technology, such as gene silencing induced in the host. Host-induced gene silencing (HIGS) technology has been applied for wheat scab resistance breeding, such as the scab resistance gene TaHRC [92], where the wild type is a pathogen-sensitive gene which produces resistance to FHB after mutation. This kind of gene can be mutated accurately in FHB-sensitive varieties through the use of gene editing technology, changing the susceptible genes into disease-resistant genes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants13081159/s1, Table S1: Previously studied effector genes of Fusarium graminearum, including their sub-cellular location, gene function, and phenotype mutation.

Author Contributions

Z.A.B.: investigation and writing—original draft; M.C., P.W. and Z.Z.: data collection; C.L. and C.Z.: investigation, N.F.A. and G.K.: data curation, methodology, formal analysis, and visualization; D.W. and K.Z.: conceptualization, resources, writing—review and editing, supervision, project administration, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Key Science and Technology Program of Henan Province (221100110300 and 2022010104) and the Shennong Laboratory (SN01-2022-01).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hao, Y.; Rasheed, A.; Zhu, Z.; Wulff, B.B.H.; He, Z. Harnessing Wheat Fhb1 for Fusarium Resistance. Trends Plant Sci. 2020, 25, 1–3. [Google Scholar] [PubMed]
  2. Stack, R.W. History of Fusarium Head Blight with Emphasis on North America. In Fusarium Head Blight of Wheat and Barley; Leonard, K.J., Bushnell, W.R., Eds.; The American Phytopathological Society: St. Paul, MI, USA, 2003. [Google Scholar]
  3. McMullen, M. Impacts of Fusarium Head Blight on the North American Agriculture Community: The Power of One Disease to Catapult Change. In Fusarium Head Blight of Wheat and Barley; Leonard, K.J., Bushnell, W.R., Eds.; The American Phytopathological Society: St. Paul, MI, USA, 2003. [Google Scholar]
  4. Burgess, L.W.; Klein, T.A.; Bryden, W.L.; Tobin, N.F. Head Blight of Wheat Caused by Fusarium-Graminearum group 1 in New South Wales Australia. Australas. Plant Pathol. 1983, 16, 72–78. [Google Scholar]
  5. Wulff, B.B.H.; Jones, J.D.G. Breeding a Fungal Gene into wheat. Science 2020, 368, 822–823. [Google Scholar] [PubMed]
  6. Parry, D.W.; Jenkinson, P.; McLeod, L. Fusarium Ear Blight (scab) in Small Grain Cereals-a Review. Plant Pathol. 1995, 44, 207–238. [Google Scholar]
  7. Jones, J.D.G.; Dangl, J.L. The plant immune system. Nature 2006, 444, 323–329. [Google Scholar] [PubMed]
  8. Schulze, S.; Yu, L.; Hua, C.; Zhang, L.; Kolb, D.; Weber, H.; Ehinger, A.; Svenja, C.; Stahl, M.; Franz-Wachtel, M.; et al. The Arabidopsis TIR-NBS-LRR Protein CSA1 Guards BAK1-BIR3 Homeostasis and Mediates Convergence of Pattern- and Effector-induced Immune responses. Cell Host Microbe 2022, 30, 1717–1731. [Google Scholar] [CrossRef] [PubMed]
  9. Zipfel, C. Plant Pattern-recognition Receptors. Trends Immunol. 2014, 35, 345–351. [Google Scholar] [PubMed]
  10. Balint-Kurti, P. The Plant Hypersensitive Response: Concepts, Control and Consequences. Mol. Plant Pathol. 2019, 20, 1163–1178. [Google Scholar] [PubMed]
  11. Blümke, A.; Falter, C.; Herrfurth, C.; Sode, B.; Bode, R.; Schäfer, W.; Feussner, I.; Voigt, C.A. Secreted fungal effector lipase releases free fatty acids to inhibit innate immunity-related callose formation during wheat head infection. Plant Physiol. 2014, 165, 346–358. [Google Scholar]
  12. Yang, C.C.; Wang, Z.Y.; Cheng, C.M. Insights into Superinfection Immunity Regulation of Xanthomonas Axonopodis Filamentous Bacteriophage cf. Curr. Microbiol. 2023, 81, 42. [Google Scholar]
  13. Petit-Houdenot, Y.; Fudal, I. Complex Interactions between Fungal Avirulence Genes and Their Corresponding Plant Resistance Genes and Consequences for Disease Resistance Management. Front. Plant Sci. 2017, 8, 1072. [Google Scholar] [PubMed]
  14. Ausubel, F.M. Are Innate Immune Signaling Pathways in Plants and Animals Conserved? Nat. Immunol. 2005, 6, 973–979. [Google Scholar] [PubMed]
  15. Flor, H.H. Current Status of the Gene-for-gene Concept. Annu. Rev. Phytopathol. 1971, 9, 275–296. [Google Scholar] [CrossRef]
  16. Flor, H.H. Inheritance of Pathogenicity in Melampsora lini. Phytopathology 1942, 32, 653–669. [Google Scholar]
  17. Trail, F. For Blighted Waves of Grain: Fusarium graminearum in the Postgenomics Era. Plant Physiol. 2009, 149, 103–110. [Google Scholar] [PubMed]
  18. Van, W.S.; Wingfield, B.D.; de Vos, L.; van der Merwe, N.A.; Santana, Q.C.; Steenkamp, E.T. Repeat-Induced Point Mutations Drive Divergence between Fusarium Circinatum and its Close Relatives. Pathogens 2019, 8, 298. [Google Scholar] [CrossRef]
  19. Klix, V.; Nowrousian, M.; Ringelberg, C.; Loros, J.J.; Dunlap, J.C.; Pöggeler, S. Functional Characterization of MAT1-1-specific Mating-type Genes in the Homothallic Ascomycete Sordaria macrospora Provides New Insights into Essential and Nonessential Sexual Regulators. Eukaryot Cell 2010, 9, 894–905. [Google Scholar]
  20. Guo, Y.; Yao, S.; Yuan, T.; Wang, Y.; Zhang, D.; Tang, W. The Spatiotemporal Control of KatG2 Catalase-peroxidase Contributes to the Invasiveness of Fusarium graminearum in Host Plants. Mol. Plant Pathol. 2019, 20, 685–700. [Google Scholar]
  21. Yun, S.H.; Arie, T.; Kaneko, S.; Yoder, O.C.; Turgeon, B.G. Molecular Organization of Mating Type Loci in Heterothallic, Homothallic and Asexual Gibberella/Fusarium species. Fungal Genet. Biol. 2000, 31, 7–20. [Google Scholar] [PubMed]
  22. Mentges, M.; Glasenapp, A.; Boenisch, M.; Malz, S.; Henrissat, B.; Frandsen, R.J.N.; Güldener, U.; Münsterkötter, M.; Bormann, J.; Lebrun, M.H.; et al. Infection Cushions of Fusarium graminearum are Fungal Arsenals for Wheat Infection. Mol. Plant Pathol. 2020, 21, 1070–1087. [Google Scholar]
  23. Ofek, P.; Yeini, E.; Arad, G.; Danilevsky, A.; Pozzi, S.; Luna, C.B.; Dangoor, S.I.; Grossman, R.; Ram, Z.; Shomron, N.; et al. Deoxyhypusine hydroxylase: A Novel Therapeutic Target Differentially Expressed in Short-term vs Long-term Survivors of Glioblastoma. Int. J. Cancer 2023, 153, 654–668. [Google Scholar]
  24. Zhao, F.; Yuan, Z.; Wen, W.; Huang, Z.; Mao, X.; Zhou, M.; Hou, Y. FgMet3 and FgMet14 Related to Cysteine and Methionine Biosynthesis Regulate Vegetative Growth, Sexual Reproduction, Rathogenicity, and Sensitivity to Fungicides in Fusarium graminearum. Front. Plant Sci. 2022, 13, 011709. [Google Scholar]
  25. Villafana, R.T.; Ramdass, A.C.; Rampersad, S.N. Selection of Fusarium Trichothecene Toxin Genes for Molecular Detection Depends on TRI Gene Cluster Organization and Gene Function. Toxins 2019, 11, 36. [Google Scholar] [CrossRef]
  26. Zhang, Y.Z.; Wei, Z.Z.; Liu, C.H.; Chen, Q.; Xu, B.J.; Guo, Z.R.; Cao, Y.L.; Wang, Y.; Han, Y.N.; Chen, C.; et al. Linoleic Acid Isomerase Gene FgLAI12 Affects Sensitivity to Salicylic Acid, Mycelial Growth and Virulence of Fusarium graminearum. Sci. Rep. 2017, 7, 46129. [Google Scholar] [CrossRef]
  27. Martinez-Rocha, A.L.; Woriedh, M.; Chemnitz, J.; Willingmann, P.; Kröger, C.; Hadeler, B.; Hauber, J.; Schäfer, W. Posttranslational Hypusination of the Eukaryotic Translation Initiation Factor-5A Regulates Fusarium graminearum Virulence. Sci. Rep. 2016, 6, 24698. [Google Scholar]
  28. Ellinger, D.; Sode, B.; Falter, C.; Voigt, C.A. Resistance of Callose Synthase Activity to Free Fatty acid Inhibition as an Indicator of Fusarium Head Blight Resistance in Wheat. Plant Signal Behav. 2014, 9, e28982. [Google Scholar]
  29. 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, 922. [Google Scholar]
  30. Kamoun, S.A. Catalogue of the Effector Secretome of Plant Pathogenic Oomycetes. Annu. Rev. Phytopathol. 2006, 44, 41–60. [Google Scholar]
  31. Kamoun, S. Groovy times: Filamentous Pathogen Effectors Revealed. Curr. Opin. Plant Biol. 2007, 10, 358–365. [Google Scholar]
  32. Zhang, X.W.; Jia, L.J.; Zhang, Y.; Jiang, G.; Li, X.; Zhang, D.; Tang, W.H. In Planta Stage-specific Fungal Gene Profiling Elucidates the Molecular Strategies of Fusarium graminearum Growing inside Wheat Coleoptiles. Plant Cell 2012, 24, 5159–5176. [Google Scholar]
  33. Kugler, K.G.; Siegwart, G.; Nussbaumer, T.; Ametz, C.; Spannagl, M.; Steiner, B.; Lemmens, M.; Mayer, K.F.; Buerstmayr, H.; Schweiger, W. Quantitative Trait Loci-dependent Analysis of a Gene Co-expression Network Associated with Fusarium Head Blight Resistance in Bread Wheat (Triticum aestivum L.). BMC Genom. 2013, 14, 728. [Google Scholar]
  34. Jiang, C.; Hei, R.; Yang, Y.; Zhang, S.; Wang, Q.; Wang, W.; Zhang, Q.; Yan, M.; Zhu, G.; Huang, P.; et al. An Orphan Protein of Fusarium graminearum Modulates Host Immunity by Mediating Proteasomal Degradation of TaSnRK1α. Nat. Commun. 2020, 11, 4382. [Google Scholar]
  35. Min, K.; Son, H.; Lee, J.; Choi, G.J.; Kim, J.C.; Lee, Y.W. Peroxisome Function Is Required for Virulence and Survival of Fusarium graminearum. Mol. Plant Microbe Interact. 2012, 25, 1617–1627. [Google Scholar]
  36. Hao, G.; McCormick, S.; Usgaard, T.; Tiley, H.; Vaughan, M.M. Characterization of Three Fusarium graminearum Effectors and Their Roles During Fusarium Head Blight. Front. Plant Sci. 2020, 11, 579553. [Google Scholar]
  37. Walton, J.D. Deconstructing the Cell Wall. Plant Physiol. 1994, 104, 1113–1118. [Google Scholar]
  38. Wang, Q.; Han, C.; Ferreira, A.O.; Yu, X.; Ye, W.; Tripathy, S.; Kale, S.D.; Gu, B.; Sheng, Y.; Sui, Y.; et al. Transcriptional Programming and Functional Interactions within the Phytophthora sojae RXLR Effector Repertoire. Plant Cell 2011, 23, 2064–2086. [Google Scholar]
  39. Brown, N.A.; Antoniw, J.; Hammond-Kosack, K.E. The Predicted Secretome of the Plant Pathogenic Fungus Fusarium graminearum: A Refined Comparative Analysis. PLoS ONE 2012, 7, e33731. [Google Scholar]
  40. Köster, P.; DeFalco, T.A.; Zipfel, C. Ca2+ Signals in Plant Immunity. EMBO J. 2022, 41, e110741. [Google Scholar]
  41. Gebrie, A. Transposable Elements as Essential Elements in the Control of Gene Expression. Mob. DNA 2023, 9, 14. [Google Scholar]
  42. Lu, S.; Edwards, M.C. Genome-wide Analysis of Small Secreted Cysteine-rich Proteins Identifies Candidate Effector Proteins Potentially Involved in Fusarium graminearum-wheat interactions. Phytopathology 2016, 106, 166–176. [Google Scholar]
  43. Lapin, G.; Van den Ackerveken, G. Susceptibility to Plant Disease: More than a Failure of Host Immunity. Trends Plant Sci. 2013, 18, 546–554. [Google Scholar]
  44. Van Schie, C.C.; Takken, F.L. Susceptibility Genes: How to be a Good Host. Annu. Rev. Phytopathol. 2014, 52, 551–581. [Google Scholar]
  45. Zaidi, S.S.; Mukhtar, M.S.; Mansoor, S. Genome Editing: Targeting Susceptibility Genes for Plant Disease Resistance. Trends Biotechnol. 2018, 36, 898–906. [Google Scholar]
  46. Guenther, J.C.; Hallen-Adams, H.E.; Bücking, H.; Shachar-Hill, Y.; Trail, F. Triacylglyceride Metabolism by Fusarium graminearum During Colonization and Sexual Development on Wheat. Mole. Plant Microbe Interact. 2009, 12, 1492–1503. [Google Scholar] [CrossRef]
  47. Hao, G.; Tiley, H.; McCormick, S. Chitin Triggers Tissue-Specific Immunity in Wheat Associated with Fusarium Head Blight. Front. Plant Sci. 2022, 13, 832502. [Google Scholar]
  48. Zuo, N.; Bai, W.Z.; Wei, W.Q.; Yuan, T.L.; Zhang, D.; Wang, Y.Z.; Tang, W.H. Fungal CFEM Effectors Negatively Regulate a Maize Wall-associated Kinase by Interacting with its Alternatively Spliced Variant to Dampen Resistance. Cell Rep. 2022, 41, 111877. [Google Scholar]
  49. Houterman, P.M.; Ma, L.; van Ooijen, G.; de Vroomen, M.J.; Cornelissen, B.J.C.; Takken, F.L.W.; Rep, M. The Effector Protein Avr2 of the Xylem-colonizing Fungus Fusarium Oxysporum Activates the Tomato Resistance Protein I-2 Intracellularly. Plant J. 2009, 58, 970–978. [Google Scholar] [CrossRef]
  50. Catanzariti, A.M.; Lim, G.T.T.; Jones, D.A. The Tomato I-3 Gene: A Novel Gene for Resistance to Fusarium wilt Disease. New Phytol. 2015, 207, 106–118. [Google Scholar] [CrossRef]
  51. Kong, X.; van Diepeningen, A.D.; van der Lee, T.A.J.; Waalwijk, C.; Xu, J.; Xu, J.; Zhang, H.; Chen, W.; Feng, J. The Fusarium graminearum Histone Acetyltransferases Are Important for Morphogenesis, DON Biosynthesis, and Pathogenicity. Front. Microbiol. 2018, 9, 654. [Google Scholar]
  52. Luger, K.; Rechsteiner, T.J.; Flaus, A.J.; Waye, M.M.Y.; Richmond, T.J. Characterization of Nucleosome Core Particles Containing Histone Proteins Made in Bacteria. J. Mol. Biol. 1997, 272, 301–311. [Google Scholar] [CrossRef]
  53. Badeaux, A.I.; Shi, Y. Emerging Roles for Chromatin as a Signal Integration and Storage Platform. Nat. Rev. Mol. Cell Biol. 2013, 14, 211–224. [Google Scholar] [CrossRef]
  54. Lanver, D.; Muller, A.N.; Happel, P.; Schweizer, G.; Haas, F.B.; Franitza, M.; Pellegrin, C.; Reissmann, S.; Altmuller, J.; Rensing, S.A.; et al. The Biotrophic Development of Ustilago Maydis Studied by RNA-Seq Analysis. Plant Cell 2018, 30, 300–323. [Google Scholar] [CrossRef]
  55. Kong, L.; Qiu, X.; Kang, J.; Wang, Y.; Chen, H.; Huang, J.; Qiu, M.; Zhao, Y.; Kong, G.; Ma, Z.; et al. A Phytophthora Effector Manipulates Host Histone Acetylation and Reprograms Defense Gene Expression to Promote Infection. Curr. Biol. 2017, 27, 981–991. [Google Scholar] [CrossRef]
  56. Ori, N.; Eshed, Y.; Paran, I.; Presting, G.; Aviv, D.; Tanksley, S.; Zamir, D.; Fluhr, R. The I2C Family from the Wilt Disease Resistance Locus I2 Belongs to the Nucleotide Binding, Leucine-rich Repeat Superfamily of Plant Resistance Genes. Plant Cell 1997, 9, 521–532. [Google Scholar]
  57. Rep, M.; Meijer, M.; Houterman, P.M.; van der Does, H.C.; Cornelissen, B.J.C. Fusarium oxysporum Evades I-3-Mediated Resistance without Altering the Matching Avirulence Gene. Mol. Plant Microbe Interact. 2005, 18, 15–23. [Google Scholar] [CrossRef]
  58. Gu, Q.; Wang, Y.; Zhao, X.; Yuan, B.; Zhang, M.; Tan, Z.; Zhang, X.; Chen, Y.; Wu, H.; Luo, Y.; et al. Inhibition of Histone Acetyltransferase GCN5 by a Transcription Factor FgPacC Controls Fungal Adaptation to Host-derived Iron Stress. Nucleic Acid Res. 2022, 50, 6190–6210. [Google Scholar] [CrossRef]
  59. Zhi, P.; Kong, L.; Liu, J.; Zhang, X.; Wang, X.; Li, H.; Sun, M.; Li, Y.; Chang, C. Histone Deacetylase TaHDT701 Functions in TaHDA6-TaHOS15 Complex to Regulate Wheat Defense Responses to Blumeria graminis f. sp. tritici. Int. J. Mol. Sci. 2020, 21, 2640. [Google Scholar] [CrossRef]
  60. Sagaram, U.S.; Pandurangi, R.; Kaur, J.; Smith, T.J.; Shah, D.M. Structure-activity Determinants in Antifungal Plant Defensins MsDef1 and MtDef4 with Different Modes of Action against Fusarium graminearum. PLoS ONE 2011, 6, e18550. [Google Scholar] [CrossRef]
  61. Fouché, S.; Oggenfuss, U.; Chanclud, E.; Croll, D.A. Devil’s Bargain with Transposable Elements in Plant Pathogens. Trends Genet. 2022, 38, 222–230. [Google Scholar] [CrossRef]
  62. Daboussi, M.J.; Capy, P. Transposable Elements in Filamentous Fungi. Annu. Rev. Microbiol. 2003, 57, 275–299. [Google Scholar] [CrossRef]
  63. Van Wyk, S.; Wingfield, B.D.; De Vos, L.; van der Merwe, N.A.; Steenkamp, N.A. Genome-wide Analysis of Repeat-Induced Point Mutations in the Ascomycota. Front. Microbiol. 2021, 11, 622368. [Google Scholar] [CrossRef]
  64. Amselem, J.; Quesneville, H.; Oliver, R.P.; Wincker, P.; Balesdent, M.H.; Howlett, B.J. Effector Diversification within Compartments of the Leptosphaeria Maculans Genome Affected by Repeat-Induced Point mutations. Nat. Commun. 2011, 2, 202. [Google Scholar]
  65. Komluski, J.; Habig, M.; Stukenbrock, E.H. Repeat-Induced Point Mutation and Gene Conversion Coinciding with Heterochromatin Shape the Genome of a Plant-Pathogenic Fungus. mBio 2023, 14, e0329022. [Google Scholar] [CrossRef]
  66. Van Wyk, S.; Harrison, C.H.; Wingfield, B.D.; De Vos, L.; van der Merwe, N.A.; Steenkamp, E.T. The RIPper, a Web-based Tool for Genome-wide Quantification of Repeat-Induced Point (RIP) Mutations. PeerJ 2019, 7, e7447. [Google Scholar] [CrossRef]
  67. Hartmann, F.E.; Sánchez-Vallet, A.; McDonald, B.A.; Croll, D.A. Fungal Wheat Pathogen Evolved Host Specialization by Extensive Chromosomal Rearrangements. ISME J. 2017, 11, 1189–1204. [Google Scholar] [CrossRef]
  68. Selker, E.U. Premeiotic Instability of Repeated Sequences in Neurospora crassa. Annu. Rev. Genet. 1990, 24, 579–613. [Google Scholar] [CrossRef]
  69. Selker, E.U. Repeat-induced Gene Silencing in Fungi. Adv. Genet. 2002, 46, 439–450. [Google Scholar]
  70. Selker, E.U.; Cambareri, E.B.; Jensen, B.C.; Haack, K. Rearrangement of Duplicated DNA in Specialized Cells of Neurospora. Cell 1987, 51, 741–752. [Google Scholar] [CrossRef]
  71. Galagan, J.E.; Selker, E.U. RIP: The Evolutionary Cost of Genome Defense. Trends Genet. 2004, 20, 417–423. [Google Scholar] [CrossRef]
  72. Son, H.; Min, K.; Lee, J.; Raju, N.B.; Lee, Y.W. Meiotic Silencing in the Homothallic Fungus Gibberella zeae. Fungal Biol. 2011, 115, 1290–1302. [Google Scholar] [CrossRef]
  73. Sridhar, P.S.; Trofimova, D.; Subramaniam, R.; González-Peña Fundora, D.; Foroud, N.A.; Allingham, J.S.; Loewen, M.C. Ste2 receptor-mediated chemotropism of Fusarium graminearum Contributes to its Pathogenicity against Wheat. Sci. Rep. 2020, 10, 10770. [Google Scholar] [CrossRef] [PubMed]
  74. Urban, M.; King, R.; Hassani-Pak, K.; Hammond-Kosack, K.E. Whole-genome Analysis of Fusarium graminearum Insertional Mutants Identifies Virulence Associated Genes and Unmasks Untagged Chromosomal Deletions. BMC Genom. 2015, 16, 261. [Google Scholar] [CrossRef] [PubMed]
  75. King, R.; Urban, M.; Hammond-Kosack, M.C.; Hassani-Pak, K.; Hammond-Kosack, K.E. The completed genome sequence of the pathogenic ascomycete fungus Fusarium graminearum. BMC Genom. 2015, 16, 544. [Google Scholar] [CrossRef] [PubMed]
  76. Kim, H.K.; Jo, S.M.; Kim, G.Y.; Kim, D.W.; Kim, Y.K.; Yun, S.H. A Large-Scale Functional Analysis of Putative Target Genes of Mating-Type Loci Provides Insight into the Regulation of Sexual Development of the Cereal Pathogen Fusarium graminearum. PLoS Genet. 2015, 11, e1005486. [Google Scholar] [CrossRef] [PubMed]
  77. Irelan, J.T.; Selker, E.U. Cytosine Methylation Associated with Repeat-induced Point Mutation Causes Epigenetic Gene Silencing in Neurospora crassa. Genetics 1997, 146, 509–523. [Google Scholar] [CrossRef]
  78. Hane, J.K.; Oliver, R.P. In silico Reversal of Repeat-induced Point Mutation (RIP) Identifies the Origins of Repeat Families and Uncovers Obscured Duplicated Genes. BMC Genom. 2010, 11, 655. [Google Scholar] [CrossRef]
  79. Chen, L.; Wang, H.; Yang, J.; Yang, X.; Zhang, M.; Zhao, Z.; Fan, Y.; Wang, C.; Wang, J. Bioinformatics and Transcriptome Analysis of CFEM Proteins in Fusarium graminearum. J. Fungi 2021, 7, 871. [Google Scholar] [CrossRef]
  80. Cai, N.; Liu, R.; Yan, D.; Zhang, N.; Zhu, K.; Zhang, D.; Nong, X.; Tu, X.; Zhang, Z.; Wang, G. Bioinformatics Analysis and Functional Characterization of the CFEM Proteins of Metarhizium anisopliae. J. Fungi 2022, 8, 661. [Google Scholar] [CrossRef]
  81. 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, 37. [Google Scholar] [CrossRef]
  82. Rocher, F.; Alouane, T.; Philippe, G.; Martin, M.L.; Label, P.; Langin, T.; Bonhomme, L. Fusarium graminearum Infection Strategy in Wheat Involves a Highly Conserved Genetic Program That Controls the Expression of a Core Effectome. Int. J. Mol. Sci. 2022, 23, 1914. [Google Scholar] [CrossRef]
  83. Hammond-Kosack, K.E.; Rudd, J.J. Plant Resistance Signalling Hijacked by a Necrotrophic Fungal Pathogen. Plant Signal. Behav. 2008, 3, 993–995. [Google Scholar] [CrossRef]
  84. Perincherry, L.; Lalak-Kańczugowska, J.; Stępień, Ł. Fusarium-Produced Mycotoxins in Plant-Pathogen Interactions. Toxins 2019, 11, 664. [Google Scholar] [CrossRef]
  85. Zhao, X.M.; Zhang, X.W.; Tang, W.H.; Chen, L. FPPI: Fusarium graminearum Protein-protein Interaction Database. J. Proteome Res. 2009, 8, 4714–4721. [Google Scholar] [CrossRef]
  86. Bahrini, I.; Ogawa, T.; Kobayashi, F.; Kawahigashi, H.; Handa, H. Overexpression of the Pathogen-inducible Wheat TaWRKY45 Gene Confers Disease Resistance to Multiple Fungi in Transgenic Wheat Plants. Breed Sci. 2011, 61, 319–326. [Google Scholar] [CrossRef]
  87. Buerstmayr, M.; Buerstmayr, H. The Effect of the Rht1 Haplotype on Fusarium head blight Resistance in Relation to Type and Level of Background Resistance and in Combination with Fhb1 and Qfhs.ifa-5A. Theor. Appl. Genet. 2022, 135, 1985–1996. [Google Scholar] [CrossRef]
  88. Mesterhazy, A. What Is Fusarium Head Blight (FHB) Resistance and What Are Its Food Safety Risks in Wheat? Problems and Solutions—A Review. Toxins 2024, 16, 31. [Google Scholar] [CrossRef]
  89. Berraies, S.; Cuthbert, R.; Knox, R.; Singh, A.; DePauw, R.; Ruan, Y.; Bokore, F.; Henriquez, M.A.; Kumar, S.; Burt, A.; et al. High-density Genetic Mapping of Fusarium head blight Resistance and Agronomic Traits in Spring Wheat. Front. Plant Sci. 2023, 14, 1134132. [Google Scholar] [CrossRef]
  90. Buerstmayr, M.; Steiner, B.; Buerstmayr, H. Breeding for Fusarium head blight Resistance in Wheat—Progress and Challenges. Plant Breed. 2020, 139, 429–454. [Google Scholar] [CrossRef]
  91. Wang, H.; Sun, S.; Ge, W.; Zhao, L.; Hou, B.; Wang, K.; Lyu, Z.; Chen, L.; Xu, S.; Guo, J.; et al. Horizontal gene transfer of Fhb7 from fungus underlies Fusarium head blight resistance in wheat. Science 2020, 368, eaba5435. [Google Scholar]
  92. Su, Z.; Bernardo, A.; Tian, B.; Chen, H.; Wang, S.; Ma, H.; Cai, S.; Liu, D.; Zhang, D.; Li, T.; et al. A deletion mutation in TaHRC confers Fhb1 resistance to Fusarium head blight in wheat. Nat Genet. 2019, 51, 1099–1105. [Google Scholar] [CrossRef]
Figure 1. Generalized life cycle of Fusarium spp. [17]. Following plasmogamy and karyogamy, out-crossed and self-perithecium produce recombinant and clone meiotic spores, respectively. These form the haploid mycelium (HM), which, in turn, form three types of mitotic spores. While conidia (micro- or macro-conidia) can colonize the host, chlamydospores, in addition to direct colonization of the crop, can overwinter and develop into perithecium to restart the cycle when conditions are favorable.
Figure 1. Generalized life cycle of Fusarium spp. [17]. Following plasmogamy and karyogamy, out-crossed and self-perithecium produce recombinant and clone meiotic spores, respectively. These form the haploid mycelium (HM), which, in turn, form three types of mitotic spores. While conidia (micro- or macro-conidia) can colonize the host, chlamydospores, in addition to direct colonization of the crop, can overwinter and develop into perithecium to restart the cycle when conditions are favorable.
Plants 13 01159 g001
Figure 2. Schematic view of effectors-based mode of invasion at different cellular locations in relation to the host response. In PTI, various signaling events occur, including activation of influx of Ca2+ into the cytosol and production of ROS [40]. Effector 1 stimulates pathogen pressure at the host cell wall (e.g., leading to degradation), Effector 2 promotes cell-to-cell proliferation, and Effector 3 suppresses the cross-talk between PTI and ETI within the cytoplasm. Effector 4 leads to host genomic re-modeling, evidenced by different expression levels of F. graminerum effectors [36].
Figure 2. Schematic view of effectors-based mode of invasion at different cellular locations in relation to the host response. In PTI, various signaling events occur, including activation of influx of Ca2+ into the cytosol and production of ROS [40]. Effector 1 stimulates pathogen pressure at the host cell wall (e.g., leading to degradation), Effector 2 promotes cell-to-cell proliferation, and Effector 3 suppresses the cross-talk between PTI and ETI within the cytoplasm. Effector 4 leads to host genomic re-modeling, evidenced by different expression levels of F. graminerum effectors [36].
Plants 13 01159 g002
Figure 3. A schematic representation of the roles of effector proteins during fungal plant interaction. Histone modifications regulate chromatin compartmentalization via phase separation and influence gene expression. (A) DNA wraps around complexes of histone proteins, active enhancers regulate gene transcription through chromatin looping with the promoters of target genes, playing a role in the regulation of gene expression. (B) The nucleosome sliding expose a region that has been previously occluded; middle, ejection of a nucleosome to expose to corresponding DNA; and bottom, the substitution of a standard nucleosome with a variant histone.
Figure 3. A schematic representation of the roles of effector proteins during fungal plant interaction. Histone modifications regulate chromatin compartmentalization via phase separation and influence gene expression. (A) DNA wraps around complexes of histone proteins, active enhancers regulate gene transcription through chromatin looping with the promoters of target genes, playing a role in the regulation of gene expression. (B) The nucleosome sliding expose a region that has been previously occluded; middle, ejection of a nucleosome to expose to corresponding DNA; and bottom, the substitution of a standard nucleosome with a variant histone.
Plants 13 01159 g003
Figure 4. Potential patterns of mutations by repeat-induced point (RIP) immediately before karyogamy of F. graminearum. This fungus differs from other filamentous fungi as it is homothallic (self-fertile) and rarely out-crosses, which limits the opportunity to acquire new repeats [18,78]. RIP identifies duplicated sequences and introduces C: G to T: A transition mutations.
Figure 4. Potential patterns of mutations by repeat-induced point (RIP) immediately before karyogamy of F. graminearum. This fungus differs from other filamentous fungi as it is homothallic (self-fertile) and rarely out-crosses, which limits the opportunity to acquire new repeats [18,78]. RIP identifies duplicated sequences and introduces C: G to T: A transition mutations.
Plants 13 01159 g004
Figure 5. Proposed model of plant resistance. Epigenetic changes within the host genome result in silencing of resistance genes against foreign invasion, which is why it is important to protect the host genome against effector interaction. This will protect the host genome, allowing it to promote its resistance mechanism against the pathogen.
Figure 5. Proposed model of plant resistance. Epigenetic changes within the host genome result in silencing of resistance genes against foreign invasion, which is why it is important to protect the host genome against effector interaction. This will protect the host genome, allowing it to promote its resistance mechanism against the pathogen.
Plants 13 01159 g005
Table 1. Fusarium graminearum name or ID of histone modification factors.
Table 1. Fusarium graminearum name or ID of histone modification factors.
Protein Name or IDGene FamilySub-Cellular LocationGo Term Annotation
FGSG_02040T0HAT_GNAT >> ELP3Nuclear, cytoplasmic  GO:0008080: N-acetyltransferase activity,
  GO:0046933: hydrogen-transporting ATP synthase activity, rotational mechanism
FGSG_00280T0HAT_GNAT >> GCN5nuclear  GO:0008080: N-acetyltransferase activity
FGSG_04254T0HAT_MYST >> Esa1nuclear  Not found
FGSG_06047T0HAT_MYST >> Sas2nuclear  Not found
FGSG_08481T0HAT_MYST >> Tip60nuclear  GO:0005515: protein binding,
  GO:0007242: intracellular signaling cascade
  GO:0006355: regulation of transcription, DNA-dependent
  GO:0008270: zinc ion binding
FGSG_02567T0HMT >> DOT1nuclear   Not found
FGSG_01134T0HMT >> PRMT_1Mitochondrial, cytoplasmic GO:0008757: S-adenosylmethionine-dependent methyltransferase activity
 GO:0005737: cytoplasm
 GO:0006479: protein amino acid methylation
 GO:0008276: protein methyltransferase activity
FGSG_10718T0HMT >> PRMT_1cytoplasmic  GO:0008757: S-adenosylmethionine-dependent methyltransferase activity
FGSG_07445T0HMT >> SET1nuclear  GO:0000166: nucleotide binding
FGSG_05558T0HMT >> SET2nuclear  Not found
FGSG_01558T0HDM >> JHDM3_JMJD2Nuclear   GO:0005515: protein binding
  GO:0006355: regulation of transcription, DNA-dependent
  GO:0008270: zinc ion binding
FGSG_00780T0HDAC >> ClassINuclear   GO:0016575: histone deacetylation
  GO:0004407: histone deacetylase activity
  GO:0005634: nucleus
FGSG_01353T0HDAC >> ClassICytoplasmic, nuclear  GO:0016575: histone deacetylation
  GO:0004407: histone deacetylase activity
  GO:0005634: nucleus
FGSG_04324T0HDAC >> ClassIIBCytoplasmic, nuclear Not found
FGSG_09218T0HDAC >> ClassIIINuclear, cytoplasmic GO:0006342: chromatin silencing
 GO:0003677: DNA binding
 GO:0006355: regulation of transcription, DNA-dependent
 GO:0005677: chromatin silencing complex
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

Buttar, Z.A.; Cheng, M.; Wei, P.; Zhang, Z.; Lv, C.; Zhu, C.; Ali, N.F.; Kang, G.; Wang, D.; Zhang, K. Update on the Basic Understanding of Fusarium graminearum Virulence Factors in Common Wheat Research. Plants 2024, 13, 1159. https://doi.org/10.3390/plants13081159

AMA Style

Buttar ZA, Cheng M, Wei P, Zhang Z, Lv C, Zhu C, Ali NF, Kang G, Wang D, Zhang K. Update on the Basic Understanding of Fusarium graminearum Virulence Factors in Common Wheat Research. Plants. 2024; 13(8):1159. https://doi.org/10.3390/plants13081159

Chicago/Turabian Style

Buttar, Zeeshan Ali, Mengquan Cheng, Panqin Wei, Ziwei Zhang, Chunlei Lv, Chenjia Zhu, Nida Fatima Ali, Guozhang Kang, Daowen Wang, and Kunpu Zhang. 2024. "Update on the Basic Understanding of Fusarium graminearum Virulence Factors in Common Wheat Research" Plants 13, no. 8: 1159. https://doi.org/10.3390/plants13081159

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