Next Article in Journal
Dissecting Source-Sink Relationship of Subtending Leaf for Yield and Fiber Quality Attributes in Upland Cotton (Gossypium hirsutum L.)
Next Article in Special Issue
Influence of Silicon on Biocontrol Strategies to Manage Biotic Stress for Crop Protection, Performance, and Improvement
Previous Article in Journal
Effect of Individual, Simultaneous and Sequential Inoculation of Pseudomonas fluorescens and Meloidogyne incognita on Growth, Biochemical, Enzymatic and Nonenzymatic Antioxidants of Tomato (Solanum lycopersicum L.)
Previous Article in Special Issue
Transgenic Expression of dsRNA Targeting the Pentalonia nigronervosa acetylcholinesterase Gene in Banana and Plantain Reduces Aphid Populations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biotechnological Resources to Increase Disease-Resistance by Improving Plant Immunity: A Sustainable Approach to Save Cereal Crop Production

by
Valentina Bigini
1,†,
Francesco Camerlengo
1,†,
Ermelinda Botticella
2,
Francesco Sestili
1,* and
Daniel V. Savatin
1,*
1
Department of Agriculture and Forest Sciences, University of Tuscia, 01100 Viterbo, Italy
2
Institute of Sciences of Food Production (ISPA), National Research Council (CNR), 73100 Lecce, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2021, 10(6), 1146; https://doi.org/10.3390/plants10061146
Submission received: 9 April 2021 / Revised: 17 May 2021 / Accepted: 29 May 2021 / Published: 4 June 2021

Abstract

:
Plant diseases are globally causing substantial losses in staple crop production, undermining the urgent goal of a 60% increase needed to meet the food demand, a task made more challenging by the climate changes. Main consequences concern the reduction of food amount and quality. Crop diseases also compromise food safety due to the presence of pesticides and/or toxins. Nowadays, biotechnology represents our best resource both for protecting crop yield and for a science-based increased sustainability in agriculture. Over the last decades, agricultural biotechnologies have made important progress based on the diffusion of new, fast and efficient technologies, offering a broad spectrum of options for understanding plant molecular mechanisms and breeding. This knowledge is accelerating the identification of key resistance traits to be rapidly and efficiently transferred and applied in crop breeding programs. This review gathers examples of how disease resistance may be implemented in cereals by exploiting a combination of basic research derived knowledge with fast and precise genetic engineering techniques. Priming and/or boosting the immune system in crops represent a sustainable, rapid and effective way to save part of the global harvest currently lost to diseases and to prevent food contamination.

1. Introduction

With the new millennium, humankind is facing issues for too long postponed. Among the greatest challenges is to meet the food demand for a rapid increase in global population, estimated to exceed 9 billion by 2050 [1]. Additionally, climate changes negatively impact crop production as well as water and land availability for agriculture [2]. Cereals are cultivated for their edible caryopses in greater quantities worldwide and provide more food energy to humans than any other crop; wheat, maize and rice are the most important crops worldwide.
Food availability and security challenge may be overcome by boosting crop yield, particularly that of cereals, and/or by reducing crop yield losses (20–40%) to pests and diseases, therefore diminishing further consequences for livelihoods, public health and the environment [3]. Usage of chemical pesticides is the most widely used method to eliminate or minimize the severity of diseases affecting crops. However, different studies are highlighting several negative side-effects of the long-term use of pesticides [4], such as toxicity in humans, effects on non-target organisms—pollinators and soil microbiota—with consequent damage to ecosystems, and pollution of water and soil systems [5]. Moreover, effectiveness of long-term use of pesticides is impeded by different levels of resistance developed by phytopathogens [6]. Crop rotation, aiming to prevent the pathogen accumulation by alternating an incompatible host, together with the introduction of plant disease resistance genes (R genes) through specific breeding programs, represents alternative methods to combat yield losses to pests. Notably, crop rotation is not always an economically viable strategy [7], whereas classical breeding programs are not applicable in some crops for which no resistant cultivars are available. In addition, pathogens can quickly overcome plant host resistance mechanisms, particularly when resistance is encoded by a single gene [8]. For example, rice cultivars that are resistant to Magnaporthe oryzae typically become ineffective every 2–3 years [9].
Due to the existing combination of these problems, food availability and safety continue to be an area of concern, with climate changes putting an ever-growing pressure on agriculture to search for further alternatives. Thus, sustainable yield increase, diminishing usage of chemicals and toxic compounds, enhancing crop resilience to biotic and abiotic stress and improving nutritional and healthiness values represent the main, concomitant, targets to be pursued in agriculture in the shortest period of time. In this scenario, it would be very difficult, if not impossible, to succeed with conventional breeding, and the role of plant sciences and biotechnology becomes crucial for the future of humankind. Therefore, to find harmless control strategies for crop disease management, we need to exploit the plant innate immunity that, if timely activated, can efficiently contrast and restrict plant infection by microorganisms. In fact, although in nature plants face many types of biotic stresses caused by various organisms including fungi, viruses, bacteria, nematodes and insects, they generally resist most pathogens, and plant infection is usually the exception, not the rule [10]. As sessile organisms, plants continuously monitor their living environments and modify, accordingly, their growth, development, and defense in order to better adapt and optimize reproductivity. Plants possess an innate ability to sense and recognize potential invading microorganisms and to mount successful defenses [10]. Only pathogens with an evolved ability to evade recognition or suppress host defense mechanisms, or both, are successful. These biotic stress agents cause different kinds of diseases, infections, and damage to cultivated plants and significantly impact crop productivity [11]. Particular attention is paid to fungal diseases, one of the most destructive groups of cereal crop pathogens and one which is favored by climate changes. They not only cause a reduction in both grain quantity and quality but can also be dangerous for human health due to the production of high concentrations of mycotoxins. Moreover, rice blast and wheat Fusarium Head Blight (FHB) or Take-all diseases can in some cases eliminate an entire cereal crop [12,13].
In this manuscript, we provide several examples of how existing biotechnological techniques can provide insights into gene function by adding, suppressing, or enhancing gene activities. Identification of key regulators involved in plant resistance/adaptation mechanisms, combined with available fast and precise biotechnological techniques, offers the potential to rapidly act on (a)biotic stress-derived yield losses, supporting crops to finally reach their full productivity in different and changing environments.

2. Plant Biotechnology: From Random to Directed, Precise and Safe Mutagenesis

Over thousands of years since 10,000 BP, humans have domesticated plants in an unconscious manner, selecting phenotypes with traits essential either for wide adaptation to different environments or improved agronomic performance. The phenotypic changes associated with adaptation under domestication pressure are referred to as “domestication syndrome” [14]. At the turn of 19th century, the introduction of Mendelian laws led to a scientific approach in crop breeding, thus representing the first revolution in the field of plant science (Figure 1).
Increased yield and abiotic and biotic resistance followed by enhanced performance in agronomical practices characterized early plant breeding programs by promoting the development of monotypic crop fields, with consequent loss of genetic variability.
The practice of hybridization followed by selection as a crop improvement strategy was initiated in the latter part of the 19th century by Vilmorin in France and by Wilhelm Rimpau in Germany in 1875 [15]. Different strategies of crossing permitted the increase of genetic variability useful to introduce desired traits in cultivars, leading to the most important modern crops [16].
One of the most important achievement that led to the green revolution was the harnessing of dwarf and semi-dwarf genes found in spontaneous or induced mutant wheats between 1950 and the late 1960s and introduced into modern cultivars by crosses [17].
Although the most common way of generating genetic variability is to mate (cross) two or more parents that have contrasting genotypes, the selection of best resulting phenotypes fostered the development of monotypic crop fields, with consequent loss of biodiversity.
Genetic variability is the basis to discover new beneficial traits and results from mutations that have occurred in genomes, either naturally or induced. Spontaneous mutations able to produce effects on phenotypes occur at low frequency in nature and the discovery of mutagenesis between 1920–1930 [18] allowed plant breeders to boost random mutation frequency by using chemical or physical mutation agents. Irradiation can cause deletions, inversion, and translocation besides point mutations, whereas the use of chemical agents strictly produces point mutations, especially transitions [19,20].
Plant breeders have used mutagenesis intensively since 1950, and to date, the FAO/IAEA Mutant Varieties Database includes more than 3300 varieties that have been released worldwide for commercial use, including more than 1500 cereal varieties.
The discovery of the DNA structure by Watson and Crick in 1953, the deciphering of genetic code in 1968, the finding of restriction enzymes in 1970 by Nathan, and the development of recombinant DNA technology in 1973 by Cohen and Boyer paved the way for the rise of modern plant biotechnology and molecular breeding (Figure 1) [21].
Some important achievements in plant sciences characterized the second half of the last century: the development of tissue culture and regeneration techniques allowing the use of embryo rescue and doubled haploid, and the genetic engineering technology including chromosome engineering and transgenesis for gene transfer between species distantly related. A further milestone in plant biotechnology was the demonstration that Agrobacterium tumefaciens Ti plasmid can be used to integrate foreign DNA into the plant genome [22,23,24]. Soon after, a direct gene transfer method, known as particle bombardment or biolistic, was established for recalcitrant monocots species, especially cereals [25,26]. Genetic manipulation quickly proved to have a great potential in functional genomics contributing to unravel essential in plant physiology mechanisms. In few years, transgenesis was widely adopted in plant breeding programs since it renders possible introgression of genes or any DNA sequence from other species and enables targeted editing of plant genome to increase genetic variability.
During 1990s, several genetic modified (GM) crops were developed and released on the market, with regulatory approvals of 44 countries reaching more than 400 GM events involving 32 different crops, among which were maize, rice and wheat.
Among the most important traits introduced in GM crops are: (i) herbicide tolerance—introduced for the first time in soybean by Monsanto—that gave rise to Roundup Ready crops; (ii) biotic and abiotic stress resistance with the introduction of a gene encoding the crystal Bt toxin, providing protection against pests, firstly in soybean and thereafter in other crops like maize; (iii) improved yield and growth; (iv) product quality (the first GM crop released on the market was the Flavr Savr tomato that slow down the fruit softening) [27]; (v) biofortification (the most prominent example is the Golden Rice, in which an entire biosynthetic pathway has been introduced into rice to produce beta-carotene in the endosperm) [28]; (vi) pharmaceuticals, in planta production of molecules and development of edible vaccines; (vii) phytoremediation [29].
The last 30 years have been full of new discoveries in the area of plant biotechnology, which are more and more used by plant scientists and breeders (Figure 1).
Starting from the 1990s, molecular markers assist in plant selection, since many genetic markers associated with loci controlling traits of agronomic interest provided the opportunity to accelerate gain from selection. The development of next-generation sequencing technologies opened the era of Genomic Selection (GS), which allowed simultaneous selection for numerous markers, estimating the effects of all together loci at phenotypic level that would otherwise have no significant effect if individually taken. GS combined with high-throughput phenotyping became a powerful tool for the selection of the best phenotype within a plant population and to discover genes associated with quantitative traits. Several resistance genes have been isolated in potato [30], wheat [31], rice [32,33], and barley [34].
In the 2000s, a new approach, the Targeting Induced Local Lesions in Genomes (TILLING) approach, was introduced to identify mutant genotypes harboring mutations in genes of interest [35]. This technique allowed researchers to obtain commercial non-transgenic, powdery-mildew-resistant bread wheat varieties [36].
In the last decades, new breeding techniques (NBTs) are rapidly emerging from advances in genomic research and for application in crop traits improvement. They enable precise, targeted, and reliable changes in the genome and do not create multiple, unknown, unintended mutations, unlike chemical or radiation-induced mutagenesis.
Genome-editing methods produce defined mutants, thus becoming a potent tool in functional genomics and crop breeding. Zinc Finger Nucleases (ZFN) and Transcription Activator-Like Effector Nucleases (TALENs) were the dominant genome editing tools until the rise of Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) and Crispr associated protein (Cas). CRISPR-Cas is an antiviral system developed by bacteria: segments of DNA containing short, repetitive base sequences (crispr RNA, crRNA) conserve the memory of intruding nucleic acids. The system is composed of genes encoding Cas nucleases, such as Cas9, and unique spacers (sequences complementary to a target genomic sequence) located in a genomic locus forming the CRISPR array along with crRNAs and trans-activating crRNA (tracrRNA), which are partially complementary to crRNAs. The RNase III processes the transcribed mRNA, releasing crRNA/tracrRNA complexes that activate and guide Cas proteins to target specific genomic loci introducing double strand breaks [37]. For the first time ever, researchers and breeders can select and target any location in the genome by the use of a short synthetic guide RNA (sgRNA) along with an endonuclease enzyme (Cas9) [38]. Due to high editing efficiency, multiplex editing capability and ease of usage, CRISPR technologies were quickly adopted for various genome-targeting purposes. For several genome-editing techniques, the resultant plants are free from foreign genes and would be indistinguishable both from plants generated by conventional breeding techniques and from naturally mutated plants. Thus, it is difficult for the plant scientific community, especially in Europe, to understand and accept the reasons why the European Court of Justice has recently restricted (almost forbidden) the infield growth of plants obtained by precision breeding techniques like CRISPR.
To date, major and minor crops, dicots and monocots, have been edited to improve traits of agronomical interest and with an increasing attention to nutritional and healthy values of derived foods [39]. Yield remains the major concern in crop breeding; the Gn1a, DEP1 and GS3 genes were edited in rice to enhance grain number and grain size [40]; knock-out mutations in wheat Grain Weight 2 (GW2) gene increases grain weight and yield [41,42]. Improvement of resistance to biotic and abiotic stresses has also been achieved through genome editing technologies. The simultaneous modification of the three homoeologs of EDR1 in wheat results in plants resistant to powdery mildew [43]; rice lines with broad-spectrum resistance to Xanthomonas have been produced by editing the promoter regions of SWEET11, SWEET13, and SWEET14 genes [44].
Food nutritional quality and safety are essential prerogatives to feed burgeoning world population and to limit malnourishment. Waltz (2016) [45] knocked out gene encoding for polyphenol oxidase (PPO), producing a non-browning mushroom; Sun et al. [46] produced high-amylose rice through targeted mutations in the SBEIIb gene; recently, DuPont Pioneer announced intentions to commercialize waxy maize obtained by knock-out of Wx1 gene [47]; the production of low immunogenic foods has been achieved by editing gliadin genes involved in celiac disease [48] and by editing α-amylase/trypsin inhibitors in wheat [49].
Genome editing techniques have also been used to accelerate the domestication of crops [50] or to create herbicide-resistant crops [51]. CRISPR-Cas technologies are constantly developing to overcome some limitations such as off-target effects, restrictive protospacer adjacent motif (PAM) sequences, and the low efficiency of homologous recombination. The discovery of new Cas9 orthologs (Cpf1, Cas13) and the introduction of prime editing by fusing Cas9 to reverse transcriptase [52] enable to extend genome editing applications. CRISPR editors represent a new genome editing approach for producing precise point mutations; nickase Cas9 (nCas9) fused to an enzyme (cytidine deaminase or adenosine deaminase) with base conversion activity, can convert one nucleotide into another [53,54]. Gene regulation can be achieved by fusing transcriptional activator or repressor to engineered Cas9 with both catalytic domains inactivated (deadCas9 also known as dCas9) and directed to specific promoter regions [55]. CRISPR offers the opportunity to edit different targets simultaneously [56] and to obtain DNA-free genome edited plants using CRISPR-Cas ribonucleoproteins (RNP) or transient expression systems to deliver DNA cassettes encoding for editing components [57]. Such technology is applied in a wide range of applications spanning from gene silencing and gene insertions to base, RNA, and epigenome editing, therefore allowing programmable editing even of the processes included in the central dogma model [58]. In light of this, researchers have now the capability to fine tune the flow of genetic information across different levels in the central dogma and to act on factors determining the epigenetic memory resulting from plant-environment interactions [59]. Thus, CRISPR represents the best way to introduce or modify genetic information to improve major and minor traits in plants. The advantages offered by CRISPR technologies (easy to adopt, efficiency, specificity) make this technique a valid substitute for any type of gene knock-out or gene insertion technique and direct the large diffusion of its applications in every area of genetic engineering. Furthermore, transgenic and RNAi lines cannot escape from being defined GM organisms, whereas CRISPR lines cannot be assimilated by these rules since the foreign DNA is not necessarily integrated into host cells to produce precise mutations. Indeed, a recently published study of the European Commission regarding the status of new genomic techniques (NGT) under Union law identified limitations to the capacity of the legislation to keep pace with scientific developments, causing implementation challenges and legal uncertainties. It concluded that the applicable legislation is not fit for the purpose of some NGTs and their products and that it needs to be adapted to scientific and technological progress. It may not be justified to apply different levels of regulatory oversight to similar products with similar levels of risk, as it is the case for plants conventionally bred and obtained from certain NGTs.

3. Increasing Disease-Resistance in Cereals by Implementing Plant Immunity Through Transgenesis

In recent years, significant efforts have been made, and results have been obtained in understanding the interplay between plants and their invaders [60]. During evolutionary warfare with pathogens, plants have evolved sophisticated detection and inducible defense systems to properly defend themselves (Figure 2). Innate immunity is the first step in defense against biotic agents and can be activated within a few minutes after pathogen sensing [61]. The faster pathogen detection occurs, the sooner proper immune responses are mounted by plants, with a consequent higher probability to restrict or block tissue invasion. Therefore, plants deploy hundreds of pattern recognition receptors (PRRs) in the cell plasma membrane, conceptually analogous to Toll-like receptors in animal cells [62], that can identify both non-self-molecules, referred to as pathogen-associated molecular patterns (PAMPs), and altered self-molecules or damage-associated molecular patterns (DAMPs) [63,64]. Ligand binding by its cognate receptor, belonging to the Receptor-Like Kinases (RLKs) or Receptor-Like Proteins (RLPs) classes, triggers the so-called PAMP/DAMP-triggered immunity (P/DTI), which includes, as major downstream signaling events, the calcium influx, a burst of reactive oxygen species (ROS), the activation of downstream signaling pathways leading to gene expression reprogramming, and the production of antimicrobic compounds [65]. A second level of the plant immune system involves plant resistance proteins able to recognize pathogen specific effectors (Avr proteins) and triggers plant defense mechanisms in a more robust way [66]. Plant intracellular immune receptors are nucleotide-binding, leucine-rich repeat receptors (NLRs), which also exist in animals [67]. This kind of resistance is called effector-triggered immunity (ETI) and often induces the hypersensitive response (HR) that includes programmed cell death in infected cells and surrounding areas [68]. Most R genes encode proteins with unique domains that contain a conserved Nucleotide Binding Site called NBS. LRR (Leucin-Rich Repeat) is the second most important domain. NB-LRR receptors may recognize pathogen effectors delivered inside the cell to favor plant colonization [69].
Traditionally, PTI and ETI have been considered to act sequentially but independently. However, recent accumulating evidence shows that the distinction between PAMPs and effectors, PRRs and R proteins, therefore between PTI and ETI, cannot strictly be maintained [70,71], suggesting an alternative model in which the two systems interact and share common elements but in which the cellular responses they evoke appear to be distinct. Analyses of specific mutants concluded that the activation of PTI is essential for ETI to function, while ETI can boost the efficiency of PTI and prolong the immune response duration.
Plant hormones, or phytohormones, are naturally occurring signaling compounds with diverse chemical properties. They play critical roles in the adaptation to environmental changes by driving proper responses, including activation of immunity, to a wide variety of biotic and abiotic stresses. The activity of a given hormone depends on its biosynthesis, conjugation, transport, and degradation as well as hormone activation and inactivation [72,73]. Although all hormones regulate several processes independently, inducible defense responses are fine-tuned by very complex crosstalk among hormone signaling outputs [74,75,76]. This enables plants both to adjust their reaction to the type of invader encountered and to efficiently use resources [77]. Interactions between hormonal activities can be either synergistic or antagonistic [78]. Such a complex and multilayered plant immune system offers different levels on which researchers could act through biotechnological approaches in order to enhance or implement plant resistance (Table 1).

3.1. Pathogen Detection

Knowledge of the plant immune system offers the opportunity to develop new strategies of intervention at the pathogen perception level (Table 1). Increased or new recognition ability may be generated in different ways, for example by intra- and interspecies introduction of PRRs from other plants with novel recognition specificity [62,83,84,100,101,102]. In a recent study, the Arabidopsis thaliana EF-Tu (elongation factor thermo unstable) receptor, abbreviated as EFR, was transferred to monocot rice to confer resistance to two Xanthomonas oryzae pv. Oryzae isolates. Rice plants expressing such receptors were able to sense the bacterial ligand of EFR and to elicit an immune response. Moreover, the EFR receptor was able to use components of the rice immune signaling pathway for its function [80]. AtEFR was also expressed in wheat [79] driven by the rice actin promoter, and the plants showed enhanced induction of defense-related genes, callose deposition, and resistance against the cereal bacterial pathogen P. syringae pv. Oryzae. In another study, a lectin receptor-like kinase gene (LecRK) of Haynaldia villosa, a diploid wheat relative, has been transferred to wheat variety Yangmai158, which is powdery mildew susceptible [93]. Transgenic wheat plants showed a significant increase in powdery mildew resistance. Moreover, dynamic changes were detected for the expression levels of ROS generating/scavenging genes and marker genes of the salicylic acid (SA) pathway.
A different original approach is represented by engineering novel recombinant PRRs by producing chimeric receptors incorporating the beneficial properties of various RLKs and RLPs [88]; important advances have been achieved, suggesting that the ectodomain of the chimera preserves ligand perception capacity, while the intracellular domain determines the output intensity [80,86,87]. Modular assemblies between Arabidopsis EFR and rice Xa21 [86] have shown that it is reliable to engineer PRRs to increase the amplitude of the induced defense response and to expand the recognition spectrum. Indeed, using the EFR-Xa21 chimera, rice Xa21 kinase domain results functional in Arabidopsis to induce signaling and quantitative immunity against the bacterium Pseudomonas syringae pv. Tomatoe and Agrobacterium tumefaciens. As rice Xa21 triggers HR-like responses, its intracellular domain has been used to generate chimeric PRR with rice OsCEPiP ectodomain [103]. The related chimera improved cell death following treatment with chitin as well as resistance to the fungal pathogen Magnaporthe oryzae [88].
Beyond pathogen-recognition strategies, a better understanding of effectors and their role has allowed interventions at the point of pathogen modulation of host responses. Identification of effector activity targets in plant, for instance, shows which host components are “manipulated” by the invaders to promote disease. In order to interfere with these components of susceptibility, this knowledge was successfully exploited by removing [104,105,106,107] or replacing them with variants that are resistant to the effector activity without losing their native function in the host [108]. For bacterial pathogens expressing transcription activator-like (TAL) effectors that activate the expression of susceptibility genes in the host, resistance can be engineered introducing deletions in the TAL DNA binding sites on the promoter of those genes [89,109]. Another approach to engineer resistance to these bacterial pathogens is to add TAL effector binding sites to a cell-death-promoting (“executor”) gene that is triggered by TAL effectors present in common pathotypes [90,93]. According to information on virulence factor/effector biology, it will be possible to select LRR proteins with new specificities, able to inhibit the growth of necrotrophic or biotrophic pathogens or to target resistance to viruses.

3.2. Boosting the Immune Signaling

P/DTI and ETI lead to the activation of the membrane-localized ion channels and an increase in the amount of cytoplasmic calcium. Other early response events include the activation of mitogen-activated protein kinases (MAPKs) [110]. Three hormones are principally involved in downstream signaling pathways caused by P/DTI and ETI: SA, jasmonic acid (JA), and ethylene (ET). Even though SA pathway stimulates resistance to biotrophic and hemibiotrophic pathogens, JA and ET pathways are typically induced upon sensing necrotrophic pathogens and chewing insects [111]. JA and SA have important roles in the activation of transcription factors controlling biotic stress responses, the interplay between different defense signaling pathways, and chemical priming to improve plant resistance through systemic acquired resistance (SAR). However, constitutive induction of SA or JA signaling, besides inducing resistance against pathogens, also leads to pleiotropic negative effects on growth and yield, a process known as growth-defense trade-off which is based on the assumption that plants can allocate resources either to growth or in defense [112]. Activated defense programs require cellular rearrangements at different levels, including machinery involved in transcription, translation, and protein secretion as well as metabolism prioritization of carbon and nitrogen towards production of defense compounds, such as pathogenesis-related (PR) proteins. Such a trade-off represents the output of a complex and fine-tuned phytohormonal crosstalk, and researchers worldwide are trying to unravel key regulatory elements to obtain resistant plants normally growing and producing. Recently, the transcription factor TL1-Binding Factor 1 (TBF1), which is quickly and transiently triggered by pathogen attacks, has been used to produce a “TBF1-cassette” consisting of an immune-inducible promoter and two pathogen-responsive upstream open-reading frames (uORFsTBF1) of the TBF1 gene. Researchers showed that the utilization of “TBF1-cassette” can enhance broad-spectrum disease resistance with minimal adverse effects on plant growth and development [91]. The timely and tissue localized induction of immunity may prevent the reduction in plant growth and yield, consequences of activated defense responses, thus overcoming the trade-off problem. Moreover, defense responses are controlled by networks of transcriptional regulators [113]. Therefore, the overexpression of specific transcription factors is a potential strategy to engineer resistance, with minimized or no effects on yield. One interesting study concern the rice gene Ideal Plant Architecture 1 (IPA1), known as OsSPL14, in which a naturally occurred allelic variant increased yield and resistance to rice blast (Table 1). Specific phosphorylation of IPA1 in response to blast infection alters IPA1 binding specificity. This change in specificity leads the protein to bind to WRKY45, a defense regulator transcription factor, and activate its expression, therefore ensuring quantitative resistance to the pathogen [93].

3.3. R Gene Transfer

Adult plant resistance (APR) or “slow rusting” wheat genes represent a class of potential transferable R genes [114]. Different APR genes are known, but only two, Lr34 and Lr67 (Table 1), have been cloned [115,116]. Lr34 encodes an ATP-binding cassette (ABC) transporter with an unknown substrate. Transgenic wheat lines expressing Lr34 gene displayed enhanced resistance to multiple biotrophic pathogens including the leaf rust pathogen and powdery mildew both at seedling and adult stages [94,117]. Similarly, the wheat Lr67 resistance gene is a specific dominant allele of a hexose transporter that provides resistance to powdery mildew and multiple rusts. Introduction of the Lr34 allele by transformation into rice [95], barley [94], sorghum [96], maize [97], and durum wheat [98] and of Lr67 into barley [99] produced resistance to a broad spectrum of biotrophic pathogens such as Puccinia triticina (wheat leaf rust), P. striiformis f. sp. Tritici (stripe rust), P. graminis f. sp. Tritici (stem rust), Blumeria graminis f. sp. Tritici (powdery mildew), P. hordei (barley leaf rust) and B. graminis f. sp. Hordei (barley powdery mildew), Magnaporthe oryzae (rice blast), P. sorghi (maize rust), and Exserohilum turcicum (northern corn leaf blight) [94,95,97]. The mechanism by which resistance is triggered by Lr34 and Lr67 is poorly understood, although it is likely that it provides the activation of biotic or abiotic stress responses allowing the host to limit pathogen development and growth.
Wheat resistance to Fusarium species has been greatly improved by expressing either a barley uridine diphosphate-dependent glucosyltransferases (UGT), HvUGT13248, involved in mycotoxin detoxification [118], or pyramided inhibitors of cell wall-degrading enzymes secreted by the fungi, such as the bean polygalacturonase inhibiting protein (PvPGIP2) and TAXI-III, a xylanase inhibitor [119]. Interestingly, greater resistance to Fusarium graminearum has been observed in wheat plants simultaneously expressing the PvPGIP2 in lemma, palea, rachis, and anthers, whereas the expression of this inhibitor only in the endosperm did not affect FHB symptom development, hinting that further spread of the pathogen in wheat tissues no longer can be blocked once it reaches the endosperm [120].

4. Increasing Disease-Resistance in Cereals by Using Gene Expression or Editing Techniques

4.1. RNA Interference (RNAi)

RNA interference (RNAi) was first discovered in plants as a molecular mechanism involved in the recognition and degradation of non-self-nucleic acids, principally directed against virus-derived sequences. In addition to its defensive role, RNAi is essential for endogenous gene expression regulation [121]. Initiation of RNAi occurs after double-stranded RNAs (dsRNAs) or endogenous microRNAs are processed by Dicer-like proteins. The resulting small interfering (si)RNAs can be recruited by Argonaute (AGO) proteins that recognize and cleave complementary strands of RNA, resulting in gene silencing. RNAi-based resistance can be engineered against many viruses by expressing “hairpin” structures, double-stranded RNA molecules that contain viral sequences, or simply by overexpressing dysfunctional viral genes [122]. Moreover, a single double-stranded RNA molecule can be processed into a variety of siRNAs and thereby effectively target several virus sequences using a single hairpin construct.
Over the last two decades, RNAi has emerged as a powerful genetic tool for scientific research. In addition to basic studies on the determination of gene function, RNA-silencing technology has been used to develop plants with increased resistance to biotic stresses (Figure 2), (Table 2) [123,124].
Indeed, the impact of RNAi technology deployed as a GM solution against viruses is clearly demonstrated in different studies [125,126,127]. Wheat dwarf virus (WDV) is a member of the Mastrevirus genus of the Geminiviridae family. This virus translates four viral proteins and causes economical losses in wheat and barley when it is transmitted to plants through leafhoppers. Kis et al. [126] targeted 13 different wheat- and barley-infecting WDV strains to identify conservative target sites and design miRNAs by using the miRNA precursor (hvu-MIR171) backbone of barley. They constructed a polycistronic artificial microRNA (amiRNA) precursor, which expresses three amiRNAs at the same time. As a result, transgenic barely plants that express amiRNAs at high levels presented no infection symptoms.
Recently, RNAi has been explored as a strategy to also control fungi and oomycetes. Fungal target genes are obvious candidates for this approach, as disruption is known to be lethal. A biotechnological method, termed host-induced gene silencing (HIGS), has emerged as a promising alternative in plant protection because it combines high selectivity for the target pathogen with minimal side effects, as compared with chemical treatments. Significant effects have been observed in transgenic Arabidopsis and barley (Hordeum vulgare) plants, expressing via HIGS a 791 nucleotide (nt) dsRNA (CYP3RNA) targeting all three CYP51 genes (FgCYP51A, FgCYP51B, FgCYP51C) of Fusarium graminearum (Fg) that led to the inhibition of fungal infection [128].
Cheng et al. [129] reported that the expression of RNAi sequences derived from an essential Fg virulence gene, the chitin synthase 3b (Chs3b), is an effective method to enhance resistance of wheat plants against fungal pathogens. Three hairpin RNAi constructs corresponding to the different regions of Chs3b were found to silence Chs3b in Fg strains. Co-expression of these three RNAi constructs in two independent elite wheat cultivar transgenic lines conferred high levels of stable and consistent resistance (combined type I and II resistance) to both Fusarium Head Blight (FHB) and Fusarium Seedling Blight (FSB).
A better understanding of this process in diverse plant-pathogen interactions may allow to better optimize HIGS strategies providing field-relevant levels of resistance [130,131,132]. In short, RNAi appears to be a promising additional control strategy in the arsenal of plant breeders against at least some pathogens. The modular nature of RNAi is especially suitable for multiplexing via synthetic biology approaches. In addition, RNAi strategies may be particularly relevant when no pathogen resistance can be identified in natural populations.

4.2. CRISPR/Cas9 Mediated Genome Editing

In plant research, NBTs are attracting a lot of attention. NBTs appear to be suitable for many different fields in plant science, such as developmental processes and adaptation/resistance to (a)biotic stresses [133]. NBTs include the most recent and powerful molecular approaches for precise genetic modifications of single or multiple gene targets. They employ site-directed nucleases to introduce double-strand breaks at predetermined sites in DNA.
The rapid increase in scientific publications documenting the use of CRISPR/Cas highlights how this technique has a greater success rate in gene modification compared to the other available nucleases. Actually, the application of CRISPR/Cas technologies to edit plant genomes is proving to be a powerful tool for future enhancement of agronomic traits in crops, qualitative and health parameters, tolerance to abiotic stress [134], and also for the improvement of biotic stress resistance (Table 2) [135].
In a recent study, MLO loci have been targeted by RNA-guided Cas9 endonuclease in bread wheat [136]. MLO encodes a protein with seven transmembrane domains localized in the plasma membrane and is ubiquitously present in monocots and dicots [36]. It had previously been reported that MLO were susceptibility genes and that homozygous loss-of-function mutants had significantly increased resistance to powdery mildew in barley, Arabidopsis, and tomato [141,142,143]. Bread wheat plants mutated by CRISPR/Cas9 in one (TaMLO-A1) of the three MLO homeoalleles showed improved resistance to Blumeria graminis f. sp. tritici infection, a finding that once again demonstrated the important role of TaMLO genes in powdery mildew disease [136]. Another example of CRISPR/Cas9-derived resistance against the same disease is the knockout of TaEDR1 [43], conferring resistance to powdery mildew in wheat. Recently, Su et al. [140] have reported that TaHRC, a gene that encodes a putative histidine-rich calcium-binding protein, is the key determinant of resistance to FHB. Authors have demonstrated that TaHRC encodes a nuclear protein conferring FHB susceptibility and that a CRISPR–Cas9-mediated deletion spanning the start codon of this gene results in FHB resistance. Plant mutants had significantly lower FHB severity than their wild type, suggesting that TaHRC affects FHB susceptibility and that loss of function of TaHRC confers Fhb1 resistance. Plants resistant to rice blast disease were generated through CRISPR/Cas9-mediated disruption of OsERF922 and OsSEC3A genes in rice [138,139]. Ossec3a mutant plants in a putative subunit of a complex involved in exocytosis revealed a pleiotropic phenotype including improved resistance against Magnaporthe oryzae, higher levels of SA and its related genes, but also dwarf stature [138]. In contrast, no alteration of different agronomic traits was observed in T1 and T2 transgene free plants mutated in the ET responsive factor (ERF) 922, a transcription factor involved in multiple stress responses. Mutant plants had a reduced number of blast lesions at both seedling and tillering stages [139].
Relatively few studies have been published on the application of the CRISPR/Cas systems to counteract crop bacterial diseases. CRISPR/Cas9 editing of OsSWEET13 has been performed in rice to achieve resistance to bacterial blight disease caused by bacterium Xanthomonas oryzae pv. oryzae [137]. OsSWEET13 is a susceptibility gene encoding a sucrose transporter involved in plant-pathogen interaction. X. oryzae produces an effector protein, PthXo2, which induces OsSWEET13 expression in the host and the consequent condition of susceptibility. Zhou et al. [137] obtained a null mutation in OsSWEET13 in order to better explore PthXo2-dependent disease susceptibility, and resultant mutants were resistant to bacterial blight. Further genome editing strategies for multiplexed recessive resistance using a combination of the major effectors and other R genes will be the next step toward achieving bacterial blight resistance.

5. Conclusions and Perspectives

Our planet is facing unprecedented challenges because of a rising and more affluent world population, while almost half of global cereal production is lost to diseases, biodiversity is diminishing at an alarming pace, and the average temperature on earth continues to rise. To meet the global challenge of food availability, we will need to shift our mentality and lifestyle, increase investments in knowledge creation, and facilitate the usage of innovative technologies, which can shorten timings to reach these goals. On the other hand, agriculture and food production must become more sustainable. The environmental footprint of agriculture needs to diminish, and farming must adapt to the rapidly changing climate. Given that we are witnessing how these aspects are threatening crop yields worldwide, all possible approaches are required to meet these challenges. Mutagenesis had, has and will certainly have an impact on crop genetics and breeding in the attempt to increase stress resistance and productivity. The ability of the DNA to mutate provides genetic variability, which is basilar for plants, as for all organisms, to evolve and adapt to environmental changes. New genotypes with new traits, likely of agronomic interest, can be created in cultivated plants by artificially inducing mutations in their genomes. Crops with improved characteristics can be obtained by both transgenesis and conventional plant breeding, respectively by adding a new gene to the genome of a crop plant or by crossing plants with desirable characteristics and selecting combinations of genes inherited from the two parent lines. With respect to when transgenesis technology was developed, our knowledge on plant gene function(s) and activity has drastically increased, satisfying one prerequisite for transgenesis to be preferable, especially in case the gene of interest do not exist in a species that can be successfully crossed with the crop. Secondly, modern high yielding crop varieties mostly result from careful selection of lines with specific combinations of value-added genes; such combination might be destroyed in the attempt to add by crossing a useful gene variant that has been newly discovered in a wild relative. Nonetheless, additional genes, closely located in the genome to the gene of interest, are almost always transferred as well. It will take several years and different generations to restore those gene combinations, even when using modern molecular breeding techniques, such as marker assisted breeding. These problems can be avoided by introducing a new added-value gene into the high yield crop by direct transformation, therefore by transgenesis. For instance, plant cultivars with R genes can be created by a transgenic approach rather than by a traditional crossing approach. An additional advantage of transgenic strategies is that linkage drag can be prevented. In resistance breeding, linkage drag refers to the undesirable reduction in crop yield and quality sometimes associated with selection of genetic resistance to disease. One of the most efficient and sustainable solutions to control plant resistance to pathogens is to use genetic modification and genome editing techniques to complement and extend modern breeding efforts.
Genome engineering techniques have made important advances over the last decades, allowing the capability not only to control but even to edit gene expression in a precise and secure manner, see Table 1 and Table 2. Genome editing allows scientists to mutate the genome of plants in a manner similar to how mutation occurs in nature, generating heritable mutations in a predictable trait-related genomic location and thus creating a series of variable phenotypes for breeding within a single generation.
The application of such biotechnological techniques in agriculture can potentially improve food availability and security by raising crop resistance to pathogens, adverse weather and soil conditions, by enhancing the adaptability of crops to different climates and by improving yields, particularly of staple food crops such as cereals (Figure 2). Biotechnology could, over the next two decades, deliver the next wave of technological change; change that could be fundamental in understanding the molecular basis of disease resistance in enough detail to make precise predictions about engineering plants to express resistance proteins that can either recognize pathogen molecules essential for pathogenicity or finely tune hormone signaling for the benefit of crop yield [144]. In this manner, it is anticipated that biotechnological approaches can engineer durable disease resistance in crops. Examples of genetic disease solutions currently available for bacterial, viral and fungal pathogens are listed in Table 1 and Table 2.
An ambitious target for the future is to continue combining science-based knowledge with biotechnological methods to develop plants that have higher resilience to (a)biotic stresses. This will enable farmers to produce high yields while decreasing the use of chemicals and water.

Author Contributions

D.V.S. and F.S. identified patterns and trends in the literature and designed the structure of the review. V.B. and F.C. prepared the tables and figures. All of the authors contributed to searching for relevant literature, carried out a critical analysis of the literature, and discussed the literature together. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Italian Ministry of Education, University, and Research (MIUR) in the frame of the initiative “Departments of excellence”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Because of space constraints, a few noteworthy publications have not been cited or discussed properly. We apologize to the authors for this lack of completeness. Parts included in Figure 2 were created with BioRender.com (accessed on 3 June 2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fróna, D.; Szenderák, J.; Harangi-Rákos, M. The Challenge of Feeding the World. Sustainability 2019, 11, 5816. [Google Scholar] [CrossRef] [Green Version]
  2. Godfray, H.C.J.; Aveyard, P.; Garnett, T.; Hall, J.W.; Key, T.J.; Lorimer, J.; Pierrehumbert, R.T.; Scarborough, P.; Springmann, M.; Jebb, S.A. Meat consumption, health, and the environment. Science 2018, 361, eaam5324. [Google Scholar] [CrossRef] [Green Version]
  3. Shiferaw, B.; Smale, M.; Braun, H.-J.; Duveiller, E.; Reynolds, M.P.; Muricho, G. Crops that feed the world 10. Past successes and future challenges to the role played by wheat in global food security. Food Secur. 2013, 5, 291–317. [Google Scholar] [CrossRef] [Green Version]
  4. Nicolopoulou-Stamati, P.; Maipas, S.; Kotampasi, C.; Stamatis, P.; Hens, L. Chemical Pesticides and Human Health: The Urgent Need for a New Concept in Agriculture. Front. Public Health 2016, 4, 148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Gill, H.K.; Garg, H. Pesticide: Environmental Impacts and Management Strategies. In Pesticides—Toxic Effects; Larramendy, M.L., Soloneski, S., Eds.; IntechOpen: Rijeka, Croatia, 2014; pp. 187–230. [Google Scholar]
  6. Hawkins, N.J.; Bass, C.; Dixon, A.; Neve, P. The evolutionary origins of pesticide resistance. Biol. Rev. 2019, 94, 135–155. [Google Scholar] [CrossRef]
  7. Bargués-Ribera, M.; Gokhale, C.S. Eco-evolutionary agriculture: Host-pathogen dynamics in crop rotations. PLoS Comput. Biol. 2020, 16, e1007546. [Google Scholar] [CrossRef] [PubMed]
  8. Poland, J.; Rutkoski, J. Advances and Challenges in Genomic Selection for Disease Resistance. Annu. Rev. Phytopathol. 2016, 54, 79–98. [Google Scholar] [CrossRef] [PubMed]
  9. Law, J.W.-F.; Ser, H.-L.; Khan, T.M.; Chuah, L.-H.; Pusparajah, P.; Chan, K.-G.; Goh, B.-H.; Lee, L.-H. The Potential of Streptomyces as Biocontrol Agents against the Rice Blast Fungus, Magnaporthe oryzae (Pyricularia oryzae). Front. Microbiol. 2017, 8, 3. [Google Scholar] [CrossRef] [Green Version]
  10. Pandey, P.; Irulappan, V.; Bagavathiannan, M.V.; Senthil-Kumar, M. Impact of Combined Abiotic and Biotic Stresses on Plant Growth and Avenues for Crop Improvement by Exploiting Physio-morphological Traits. Front. Plant Sci. 2017, 8, 537. [Google Scholar] [CrossRef] [Green Version]
  11. Gull, A.; Lone, A.A.; Wani, N.U.I. Biotic and Abiotic Stresses in Plants. Abiotic Biot. Stress Plants 2019. [Google Scholar] [CrossRef] [Green Version]
  12. Cook, R.J. Advances in Plant Health Management in the Twentieth Century. Annu. Rev. Phytopathol. 2000, 38, 95–116. [Google Scholar] [CrossRef]
  13. Hernández-Restrepo, M.; Groenewald, J.; Elliott, M.; Canning, G.; McMillan, V.; Crous, P. Take-all or nothing. Stud. Mycol. 2016, 83, 19–48. [Google Scholar] [CrossRef] [Green Version]
  14. Meyer, R.S.; Purugganan, M.D. Evolution of crop species: Genetics of domestication and diversification. Nat. Rev. Genet. 2013, 14, 840–852. [Google Scholar] [CrossRef]
  15. Bonjean, A.P.; Angus, W.J.; Sági, F. The World Wheat Book: A History of Wheat Breeding. Cereal Res. Commun. 2001, 29, 459. [Google Scholar] [CrossRef] [Green Version]
  16. Briggs, F.N. The Use of the Backcross in Crop Improvement. Am. Nat. 1938, 72, 285–292. [Google Scholar] [CrossRef]
  17. Borlaug, N.E. Contributions of Conventional Plant Breeding to Food Production. Science 1983, 219, 689–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Stadler, L.J. Genetic Effects of X-Rays in Maize. Proc. Natl. Acad. Sci. USA 1928, 14, 69–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Leitao, J.M. Chemical mutagenesis. In Plant Mutation Breeding and Biotechnology; CABI Publishing: Vienna, Austria, 2012; pp. 135–158. [Google Scholar]
  20. Mba, C.; Afza, R.; Shu, Q.Y. Mutagenic Radiations: X-Rays, Ionizing Particles and Ul-traviolet. Plant Mutat. Breed. Biotechnol. 2012, 83–90. [Google Scholar] [CrossRef]
  21. Bhatia, S.; History, Scope and Development of Biotechnology. Introduction to Pharmaceutical Biotechnology. 2018. Available online: https://iopscience.iop.org/chapter/978-0-7503-1299-8/bk978-0-7503-1299-8ch1.pdf (accessed on 3 June 2021).
  22. Zambryski, P.; Depicker, A.; Kruger, K.; Goodman, H.M. Tumor Induction by Agro-bacterium Tumefaciens: Analysis of the Boundaries of T-DNA. J. Mol. Appl. Genet. 1982, 1, 361–370. [Google Scholar] [PubMed]
  23. Herrera-Estrella, L.; Depicker, A.; Van Montagu, M.; Schell, J. Expression of chimaeric genes transferred into plant cells using a Ti-plasmid-derived vector. Nat. Cell Biol. 1983, 303, 209–213. [Google Scholar] [CrossRef]
  24. De Block, M.; Herrera-Estrella, L.; Van Montagu, M.; Schell, J.; Zambryski, P. Expression of foreign genes in regenerated plants and in their progeny. EMBO J. 1984, 3, 1681–1689. [Google Scholar] [CrossRef] [PubMed]
  25. Klein, T.M.; Wolf, E.D.; Wu, R.; Sanford, J.C. High-velocity microprojectiles for delivering nucleic acids into living cells. Nat. Cell Biol. 1987, 327, 70–73. [Google Scholar] [CrossRef]
  26. Vasil, V.; Castillo, A.M.; Fromm, M.E.; Vasil, I.K. Herbicide Resistant Fertile Transgenic Wheat Plants Obtained by Microprojectile Bombardment of Regenerable Embryogenic Callus. Nat. Biotechnol. 1992, 10, 667–674. [Google Scholar] [CrossRef] [Green Version]
  27. Kramer, M.G.; Redenbaugh, K. Commercialization of a tomato with an antisense polygalacturonase gene: The FLAVR SAVR? tomato story. Euphytica 1994, 79, 293–297. [Google Scholar] [CrossRef]
  28. Ye, X.; Al-Babili, S.; Klöti, A.; Zhang, J.; Lucca, P.; Beyer, P.; Potrykus, I. Engineering the Provitamin A (-Carotene) Biosynthetic Pathway into (Carotenoid-Free) Rice Endosperm. Science 2000, 287, 303–305. [Google Scholar] [CrossRef] [Green Version]
  29. Bizily, S.P.; Rugh, C.L.; Meagher, R.B. Phytodetoxification of hazardous organomercurials by genetically engineered plants. Nat. Biotechnol. 2000, 18, 213–217. [Google Scholar] [CrossRef]
  30. Wang, M.; Allefs, S.; Berg, R.G.V.D.; Vleeshouwers, V.G.A.A.; Van Der Vossen, E.A.G.; Vosman, B. Allele mining in Solanum: Conserved homologues of Rpi-blb1 are identified in Solanum stoloniferum. Theor. Appl. Genet. 2008, 116, 933–943. [Google Scholar] [CrossRef]
  31. Bhullar, N.K.; Zhang, Z.; Wicker, T.; Keller, B. Wheat gene bank accessions as a source of new alleles of the powdery mildew resistance gene Pm3: A large scale allele mining project. BMC Plant Biol. 2010, 10, 88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Ramkumar, G.; Srinivasarao, K.; Mohan, K.M.; Sudarshan, I.; Sivaranjani, A.K.P.; Gopalakrishna, K.; Neeraja, C.N.; Balachandran, S.M.; Sundaram, R.M.; Prasad, M.S.; et al. Development and validation of functional marker targeting an InDel in the major rice blast disease resistance gene Pi54 (Pik h). Mol. Breed. 2010, 27, 129–135. [Google Scholar] [CrossRef]
  33. Wang, D.; Guo, C.; Huang, J.; Yang, S.; Tian, D.; Zhang, X. Allele-mining of rice blast resistance genes at AC134922 locus. Biochem. Biophys. Res. Commun. 2014, 446, 1085–1090. [Google Scholar] [CrossRef] [PubMed]
  34. Biselli, C.; Urso, S.; Tacconi, G.; Steuernagel, B.; Schulte, D.; Gianinetti, A.; Bagnaresi, P.; Stein, N.; Cattivelli, L.; Valè, G. Haplotype variability and identification of new functional alleles at the Rdg2a leaf stripe resistance gene locus. Theor. Appl. Genet. 2013, 126, 1575–1586. [Google Scholar] [CrossRef]
  35. McCallum, C.M.; Comai, L.; Greene, E.A.; Henikoff, S. Targeting Induced LocalLesions IN Genomes (TILLING) for Plant Functional Genomics. Plant Physiol. 2000, 123, 439–442. [Google Scholar] [CrossRef] [Green Version]
  36. Acevedo-Garcia, J.; Spencer, D.; Thieron, H.; Reinstädler, A.; Hammond-Kosack, K.; Phillips, A.L.; Panstruga, R. mlo-based powdery mildew resistance in hexaploid bread wheat generated by a non-transgenic TILLING approach. Plant Biotechnol. J. 2016, 15, 367–378. [Google Scholar] [CrossRef] [PubMed]
  37. Hille, F.; Richter, H.; Wong, S.P.; Bratovič, M.; Ressel, S.; Charpentier, E. The Biology of CRISPR-Cas: Backward and Forward. Cell 2018, 172, 1239–1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A Programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  39. De Sousa, C.S.; Mendonça, M.A.C.; Hassan, S.S.; Barh, D.; De Carvalho Azevedo, V.A. Biotechnology for improved crop productivity and quality. Appl. Mol. Biotechnol. 2016, 231. [Google Scholar] [CrossRef]
  40. Li, M.; Li, X.; Zhou, Z.; Wu, P.; Fang, M.; Pan, X.; Lin, Q.; Luo, W.; Wu, G.; Li, H. Reassessment of the Four Yield-related Genes Gn1a, DEP1, GS3, and IPA1 in Rice Using a CRISPR/Cas9 System. Front. Plant Sci. 2016, 7, 377. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, Y.; Li, D.; Zhang, D.; Zhao, X.; Cao, X.; Dong, L.; Liu, J.; Chen, K.; Zhang, H.; Gao, C.; et al. Analysis of the functions ofTaGW2homoeologs in wheat grain weight and protein content traits. Plant J. 2018, 94, 857–866. [Google Scholar] [CrossRef] [Green Version]
  42. Sestili, F.; Pagliarello, R.; Zega, A.; Saletti, R.; Pucci, A.; Botticella, E.; Masci, S.; Tundo, S.; Moscetti, I.; Foti, S.; et al. Enhancing grain size in durum wheat using RNAi to knockdown GW2 genes. Theor. Appl. Genet. 2018, 132, 419–429. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Bai, Y.; Wu, G.; Zou, S.; Chen, Y.; Gao, C.; Tang, D. Simultaneous modification of three homoeologs of TaEDR1 by genome editing enhances powdery mildew resistance in wheat. Plant J. 2017, 91, 714–724. [Google Scholar] [CrossRef] [Green Version]
  44. Xu, Z.; Xu, X.; Gong, Q.; Li, Z.; Li, Y.; Wang, S.; Yang, Y.; Ma, W.; Liu, L.; Zhu, B.; et al. Engineering Broad-Spectrum Bacterial Blight Resistance by Simultaneously Disrupting Variable TALE-Binding Elements of Multiple Susceptibility Genes in Rice. Mol. Plant 2019, 12, 1434–1446. [Google Scholar] [CrossRef] [Green Version]
  45. Waltz, E. Gene-edited CRISPR mushroom escapes US regulation. Nat. Cell Biol. 2016, 532, 293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Sun, Y.; Jiao, G.; Liu, Z.; Zhang, X.; Li, J.; Guo, X.; Du, W.; Du, J.; Francis, F.; Zhao, Y.; et al. Generation of High-Amylose Rice through CRISPR/Cas9-Mediated Targeted Mutagenesis of Starch Branching Enzymes. Front. Plant Sci. 2017, 8, 298. [Google Scholar] [CrossRef] [PubMed]
  47. Kerr-Enskat, K. DuPont Pioneer Announces Intentions to Commercialize First CRISPR-Cas Product. Press Release. 2016. Available online: https://www.prweb.com/releases/dupont-pioneer-seed/crispr-cas-corn/prweb13349828.htm (accessed on 1 June 2021).
  48. León, S.S.; Gil-Humanes, J.; Ozuna, C.V.; Giménez, M.J.; Sousa, C.; Voytas, D.; Barro, F. Low-gluten, nontransgenic wheat engineered with CRISPR/Cas9. Plant Biotechnol. J. 2017, 16, 902–910. [Google Scholar] [CrossRef] [PubMed]
  49. Camerlengo, F.; Frittelli, A.; Sparks, C.; Doherty, A.; Martignago, D.; Larré, C.; Lupi, R.; Sestili, F.; Masci, S. CRISPR-Cas9 Multiplex Editing of the α-Amylase/Trypsin Inhibitor Genes to Reduce Allergen Proteins in Durum Wheat. Front. Sustain. Food Syst. 2020, 4, 104. [Google Scholar] [CrossRef]
  50. Li, T.; Yang, X.; Yu, Y.; Si, X.; Zhai, X.; Zhang, H.; Dong, W.; Gao, C.; Xu, C. Domestication of wild tomato is accelerated by genome editing. Nat. Biotechnol. 2018, 36, 1160–1163. [Google Scholar] [CrossRef]
  51. Sun, Y.; Zhang, X.; Wu, C.; He, Y.; Ma, Y.; Hou, H.; Guo, X.; Du, W.; Zhao, Y.; Xia, L. Engineering Herbicide-Resistant Rice Plants through CRISPR/Cas9-Mediated Homologous Recombination of Acetolactate Synthase. Mol. Plant 2016, 9, 628–631. [Google Scholar] [CrossRef] [Green Version]
  52. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nat. Cell Biol. 2019, 576, 149–157. [Google Scholar] [CrossRef]
  53. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [Green Version]
  54. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 2017, 551, 464–471. [Google Scholar] [CrossRef]
  55. Qi, L.S.; Larson, M.H.; Gilbert, L.A.; Doudna, J.A.; Weissman, J.S.; Arkin, A.P.; Lim, W.A. Repurposing CRISPR as an RNA-Guided Platform for Sequence-Specific Control of Gene Expression. Cell 2013, 152, 1173–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Xie, K.; Minkenberg, B.; Yang, Y. Boosting CRISPR/Cas9 multiplex editing capability with the endogenous tRNA-processing system. Proc. Natl. Acad. Sci. USA 2015, 112, 3570–3575. [Google Scholar] [CrossRef] [Green Version]
  57. Woo, J.W.; Kim, J.; Kwon, S.I.; Corvalán, C.; Cho, S.W.; Kim, H.; Kim, S.-G.; Kim, S.-T.; Choe, S.; Kim, J.-S. DNA-free genome editing in plants with preassembled CRISPR-Cas9 ribonucleoproteins. Nat. Biotechnol. 2015, 33, 1162–1164. [Google Scholar] [CrossRef]
  58. Crick, F. Central Dogma of Molecular Biology. Nat. Cell Biol. 1970, 227, 561–563. [Google Scholar] [CrossRef] [PubMed]
  59. Pramanik, D.; Shelake, R.M.; Kim, M.J.; Kim, J.-Y. CRISPR-Mediated Engineering across the Central Dogma in Plant Biology for Basic Research and Crop Improvement. Mol. Plant 2021, 14, 127–150. [Google Scholar] [CrossRef] [PubMed]
  60. Roychowdhury, R.; Choudhury, S.; Hasanuzzaman, M.; Srivastava, S. Sustainable Agriculture in the Era of Climate Change; Springer Science and Business Media LLC: Berlin, Germany, 2020. [Google Scholar] [CrossRef]
  61. Malik, N.A.A.; Kumar, I.S.; Nadarajah, K. Elicitor and Receptor Molecules: Orchestrators of Plant Defense and Immunity. Int. J. Mol. Sci. 2020, 21, 963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Kawashima, C.G.; Guimarães, G.A.; Nogueira, S.R.; MacLean, D.; Cook, D.R.; Steuernagel, B.; Baek, J.; Bouyioukos, C.; Melo, B.D.V.A.; Tristão, G.; et al. A pigeonpea gene confers resistance to Asian soybean rust in soybean. Nat. Biotechnol. 2016, 34, 661–665. [Google Scholar] [CrossRef]
  63. Claus, L.A.N.; Savatin, D.V.; Russinova, E. The crossroads of receptor-mediated signaling and endocytosis in plants. J. Integr. Plant Biol. 2018, 60, 827–840. [Google Scholar] [CrossRef] [Green Version]
  64. Savatin, D.V.; Gramegna, G.; Modesti, V.; Cervone, F. Wounding in the plant tissue: The defense of a dangerous passage. Front. Plant Sci. 2014, 5, 470. [Google Scholar] [CrossRef] [Green Version]
  65. He, Y.; Zhou, J.; Shan, L.; Meng, X. Plant cell surface receptor-mediated signaling—A common theme amid diversity. J. Cell Sci. 2018, 131, jcs209353. [Google Scholar] [CrossRef] [Green Version]
  66. Spoel, S.H.; Dong, X. How do plants achieve immunity? Defence without specialized immune cells. Nat. Rev. Immunol. 2012, 12, 89–100. [Google Scholar] [CrossRef]
  67. Jones, J.D.G.; Vance, R.E.; Dangl, J.L. Intracellular innate immune surveillance devices in plants and animals. Science 2016, 354, aaf6395. [Google Scholar] [CrossRef] [Green Version]
  68. Mur, L.A.J.; Kenton, P.; Lloyd, A.J.; Ougham, H.; Prats, E. The hypersensitive response; the centenary is upon us but how much do we know? J. Exp. Bot. 2007, 59, 501–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Dodds, P.N.; Rathjen, J. Plant immunity: Towards an integrated view of plant–pathogen interactions. Nat. Rev. Genet. 2010, 11, 539–548. [Google Scholar] [CrossRef] [PubMed]
  70. Yuan, M.; Jiang, Z.; Bi, G.; Nomura, K.; Liu, M.; He, S.Y.; Zhou, J.M.; Xin, X.F. Pat-tern-Recognition Receptors Are Required for NLR-Mediated Plant Immunity. Nature 2020, 592, 105–109. [Google Scholar] [CrossRef] [Green Version]
  71. Ngou, B.P.M.; Ahn, H.-K.; Ding, P.; Jones, J.D.G. Mutual potentiation of plant immunity by cell-surface and intracellular receptors. Nat. Cell Biol. 2021, 592, 1–6. [Google Scholar] [CrossRef]
  72. Weyers, J.D.B.; Paterson, N.W. Plant hormones and the control of physiological processes. New Phytol. 2001, 152, 375–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Del Bianco, M.; Giustini, L.; Sabatini, S. Spatiotemporal changes in the role of cytokinin during root development. New Phytol. 2013, 199, 324–338. [Google Scholar] [CrossRef] [PubMed]
  74. Depuydt, S.; Hardtke, C.S. Hormone Signalling Crosstalk in Plant Growth Regulation. Curr. Biol. 2011, 21, R365–R373. [Google Scholar] [CrossRef]
  75. Vanstraelen, M.; Benková, E. Hormonal Interactions in the Regulation of Plant Development. Annu. Rev. Cell Dev. Biol. 2012, 28, 463–487. [Google Scholar] [CrossRef]
  76. Bargmann, B.; Vanneste, S.; Krouk, G.; Nawy, T.; Efroni, I.; Shani, E.; Choe, G.; Friml, J.; Bergmann, D.C.; Estelle, M.; et al. A map of cell type-specific auxin responses. Mol. Syst. Biol. 2013, 9, 688. [Google Scholar] [CrossRef] [Green Version]
  77. Pieterse, C.M.; Van Der Does, D.; Zamioudis, C.; Leon-Reyes, A.; Van Wees, S. Hormonal Modulation of Plant Immunity. Annu. Rev. Cell Dev. Biol. 2012, 28, 489–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Garay-Arroyo, A.; Sánchez, M.D.L.P.; García-Ponce, B.; Azpeitia, E.; Álvarez-Buylla, E.R. Hormone symphony during root growth and development. Dev. Dyn. 2012, 241, 1867–1885. [Google Scholar] [CrossRef] [PubMed]
  79. Schoonbeek, H.; Wang, H.; Stefanato, F.L.; Craze, M.; Bowden, S.; Wallington, E.; Zipfel, C.; Ridout, C.J. Arabidopsis EF -Tu receptor enhances bacterial disease resistance in transgenic wheat. New Phytol. 2015, 206, 606–613. [Google Scholar] [CrossRef]
  80. Schwessinger, B.; Bahar, O.; Thomas, N.; Holton, N.; Nekrasov, V.; Ruan, D.; Canlas, P.E.; Daudi, A.; Petzold, C.; Singan, V.R.; et al. Transgenic Expression of the Dicotyledonous Pattern Recognition Receptor EFR in Rice Leads to Ligand-Dependent Activation of Defense Responses. PLoS Pathog. 2015, 11, e1004809. [Google Scholar] [CrossRef] [Green Version]
  81. Lu, F.; Wang, H.; Wang, S.; Jiang, W.; Shan, C.; Li, B.; Yang, J.; Zhang, S.; Sun, W. Enhancement of innate immune system in monocot rice by transferring the dicotyledonous elongation factor Tu receptor EFR. J. Integr. Plant Biol. 2014, 57, 641–652. [Google Scholar] [CrossRef] [PubMed]
  82. Peng, H.; Chen, Z.; Fang, Z.; Zhou, J.; Xia, Z.; Gao, L.; Chen, L.; Li, L.; Li, T.; Zhai, W.; et al. Rice Xa21 primed genes and pathways that are critical for combating bacterial blight infec-tion. Sci Rep. 2015, 5, 12165. [Google Scholar] [CrossRef] [Green Version]
  83. Chen, T.; Xiao, J.; Xu, J.; Wan, W.; Qin, B.; Cao, A.; Chen, W.; Xing, L.; Du, C.; Gao, X.; et al. Two members of TaRLK family confer powdery mildew resistance in common wheat. BMC Plant Biol. 2016, 16, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Rajaraman, J.; Douchkov, D.; Hensel, G.; Stefanato, F.L.; Gordon, A.; Ereful, N.; Caldararu, O.F.; Petrescu, A.-J.; Kumlehn, J.; Boyd, L.A.; et al. An LRR/Malectin Receptor-Like Kinase Mediates Resistance to Non-adapted and Adapted Powdery Mildew Fungi in Barley and Wheat. Front. Plant Sci. 2016, 7, 1836. [Google Scholar] [CrossRef] [Green Version]
  85. Wang, Z.; Cheng, J.; Fan, A.; Zhao, J.; Yu, Z.; Li, Y.; Wang, X. LecRK-V, an L-type lectin receptor kinase in Haynaldia villosa, plays positive role in resistance to wheat powdery mildew. Plant Biotechnol. J. 2018, 16, 50–62. [Google Scholar] [CrossRef] [Green Version]
  86. Holton, N.; Nekrasov, V.; Ronald, P.C.; Zipfel, C. The Phylogenetically-Related Pattern Recognition Receptors EFR and XA21 Recruit Similar Immune Signaling Components in Monocots and Dicots. PLoS Pathog. 2015, 11, e1004602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Thomas, N.C.; Oksenberg, N.; Liu, F.; Caddell, D.; Nalyvayko, A.; Nguyen, Y.; Schwessinger, B.; Ronald, P.C. The rice XA21 ectodomain fused to the Arabidopsis EFR cytoplasmic domain confers resistance to Xanthomonas oryzae pv. oryzae. PeerJ 2018, 6, e4456. [Google Scholar] [CrossRef] [Green Version]
  88. Boutrot, F.; Zipfel, C. Function, Discovery, and Exploitation of Plant Pattern Recognition Receptors for Broad-Spectrum Disease Resistance. Annu. Rev. Phytopathol. 2017, 55, 257–286. [Google Scholar] [CrossRef] [PubMed]
  89. Li, T.; Liu, B.; Spalding, M.; Weeks, D.; Yang,, B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nat. Biotechnol. 2012, 30, 390–392. [Google Scholar] [CrossRef] [PubMed]
  90. Hummel, A.W.; Doyle, E.L.; Bogdanove, A.J. Addition of transcription activator-like effector binding sites to a pathogen strain-specific rice bacterial blight resistance gene makes it effective against additional strains and against bacterial leaf streak. New Phytol. 2012, 195, 883–893. [Google Scholar] [CrossRef]
  91. Xu, G.; Yuan, M.; Ai, C.; Liu, L.; Zhuang, E.; Karapetyan, S.; Wang, S.; Dong, X. uORF-mediated translation allows engineered plant disease resistance without fitness costs. Nat. Cell Biol. 2017, 545, 491–494. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, Z.; Liu, X.; Wang, X.; Zhou, M.; Zhou, X.; Ye, X.; Wei, X. An R2R3 MYB transcription factor in wheat, TaPIMP1, mediates host resistance to Bipolaris sorokiniana and drought stresses through regulation of defense- and stress-related genes. New Phytol. 2012, 196, 1155–1170. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, J.; Zhou, L.; Shi, H.; Chern, M.; Yu, H.; Yi, H.; He, M.; Yin, J.; Zhu, X.; Li, Y.; et al. A single transcription factor promotes both yield and immunity in rice. Science 2018, 361, 1026–1028. [Google Scholar] [CrossRef] [Green Version]
  94. Risk, J.M.; Selter, L.L.; Chauhan, H.; Krattinger, S.G.; Kumlehn, J.; Hensel, G.; Viccars, L.A.; Richardson, T.M.; Buesing, G.; Troller, A.; et al. The wheat Lr34 gene provides resistance against multiple fungal pathogens in barley. Plant Biotechnol. J. 2013, 11, 847–854. [Google Scholar] [CrossRef]
  95. Krattinger, S.G.; Sucher, J.; Selter, L.L.; Chauhan, H.; Zhou, B.; Tang, M.; Upadhyaya, N.M.; Mieulet, D.; Guiderdoni, E.; Weidenbach, D.; et al. The wheat durable, multipathogen resistance geneLr34confers partial blast resistance in rice. Plant Biotechnol. J. 2016, 14, 1261–1268. [Google Scholar] [CrossRef] [Green Version]
  96. Schnippenkoetter, W.; Lo, C.; Liu, G.; Dibley, K.; Chan, W.L.; White, J.; Milne, R.; Zwart, A.; Kwong, E.; Keller, B.; et al. The wheat Lr34 multipathogen resistance gene confers resistance to anthracnose and rust in sorghum. Plant Biotechnol. J. 2017, 15, 1387–1396. [Google Scholar] [CrossRef]
  97. Sucher, J.; Boni, R.; Yang, P.; Rogowsky, P.; Büchner, H.; Kastner, C.; Kumlehn, J.; Krattinger, S.G.; Keller, B. The durable wheat disease resistance geneLr34confers common rust and northern corn leaf blight resistance in maize. Plant Biotechnol. J. 2016, 15, 489–496. [Google Scholar] [CrossRef]
  98. Rinaldo, A.; Gilbert, B.; Boni, R.; Krattinger, S.G.; Singh, D.; Park, R.F.; Lagudah, E.; Ayliffe, M. TheLr34adult plant rust resistance gene provides seedling resistance in durum wheat without senescence. Plant Biotechnol. J. 2017, 15, 894–905. [Google Scholar] [CrossRef] [Green Version]
  99. Milne, R.J.; Dibley, K.E.; Schnippenkoetter, W.; Mascher, M.; Lui, A.C.; Wang, L.; Lo, C.; Ashton, A.R.; Ryan, P.R.; Lagudah, E.S. The Wheat Lr67 Gene from the Sugar Transport Protein 13 Family Confers Multipathogen Resistance in Barley. Plant Physiol. 2019, 179, 1285–1297. [Google Scholar] [CrossRef] [Green Version]
  100. Albert, I.; Böhm, H.; Albert, M.; Feiler, C.E.; Imkampe, J.; Wallmeroth, N.; Brancato, C.; Raaymakers, T.M.; Oome, S.; Zhang, H.; et al. An RLP23–SOBIR1–BAK1 complex mediates NLP-triggered immunity. Nat. Plants 2015, 1, 15140. [Google Scholar] [CrossRef]
  101. Steuernagel, B.; Periyannan, S.K.; Hernández-Pinzón, I.; Witek, K.; Rouse, M.N.; Yu, G.; Hatta, A.; Ayliffe, M.; Bariana, H.; Jones, J.D.G.; et al. Rapid cloning of disease-resistance genes in plants using mutagenesis and sequence capture. Nat. Biotechnol. 2016, 34, 652–655. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, W.; Pan, Q.; He, F.; Akhunova, A.; Chao, S.; Trick, H.; Akhunov, E. Transgenerational CRISPR-Cas9 Activity Facilitates Multiplex Gene Editing in Allopolyploid Wheat. CRISPR J. 2018, 1, 65–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Kouzai, Y.; Kaku, H.; Shibuya, N.; Minami, E.; Nishizawa, Y. Expression of the chimeric receptor between the chitin elicitor receptor CEBiP and the receptor-like protein kinase Pi-d2 leads to enhanced responses to the chitin elicitor and disease resistance against Magnaporthe oryzae in rice. Plant Mol. Biol. 2012, 81, 287–295. [Google Scholar] [CrossRef]
  104. Bozkurt, T.O.; Richardson, A.; Dagdas, Y.F.; Mongrand, S.; Kamoun, S.; Raffaele, S. The Plant Membrane-Associated REMORIN1.3 Accumulates in Discrete Perihaustorial Domains and Enhances Susceptibility to Phytophthora infestans. Plant Physiol. 2014, 165, 1005–1018. [Google Scholar] [CrossRef] [Green Version]
  105. Boevink, P.C.; McLellan, H.; Gilroy, E.M.; Naqvi, S.; He, Q.; Yang, L.; Wang, X.; Turnbull, D.; Armstrong, M.R.; Tian, Z.; et al. Oomycetes Seek Help from the Plant: Phytophthora infestans Effectors Target Host Susceptibility Factors. Mol. Plant 2016, 9, 636–638. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, B.-H.; Hagemann, S.; Mamareli, P.; Lauer, U.M.; Hoffmann, U.; Beckstette, M.; Fohse, L.; Prinz, I.; Pezoldt, J.; Suerbaum, S.; et al. Foxp3+ T cells expressing RORγt represent a stable regulatory T-cell effector lineage with enhanced suppressive capacity during intestinal inflammation. Mucosal Immunol. 2016, 9, 444–457. [Google Scholar] [CrossRef] [PubMed]
  107. Murphy, F.; He, Q.; Armstrong, M.; Giuliani, L.M.; Boevink, P.C.; Zhang, W.; Tian, Z.; Birch, P.R.J.; Gilroy, E.M. The Potato MAP3K StVIK Is Required for the Phytophthora infestans RXLR Effector Pi17316 to Promote Disease. Plant Physiol. 2018, 177, 398–410. [Google Scholar] [CrossRef] [Green Version]
  108. Ezhang, J.; Eyin, Z.; Ewhite, F. TAL effectors and the executor R genes. Front. Plant Sci. 2015, 6, 641. [Google Scholar] [CrossRef]
  109. Jia, H.; Zhang, Y.; Orbović, V.; Xu, J.; White, F.F.; Jones, J.B.; Wang, N. Genome editing of the disease susceptibility gene CsLOB1 in citrus confers resistance to citrus canker. Plant Biotechnol. J. 2017, 15, 817–823. [Google Scholar] [CrossRef] [Green Version]
  110. Muthamilarasan, M.; Prasad, M. Plant innate immunity: An updated insight into defense mechanism. J. Biosci. 2013, 38, 433–449. [Google Scholar] [CrossRef]
  111. Vleesschauwer, D.E.; Exu, J.; Hãfte, M. Making sense of hormone-mediated defense networking: From rice to Arabidopsis. Front. Plant Sci. 2014, 5, 611. [Google Scholar] [CrossRef] [PubMed]
  112. Figueroa-Macías, J.P.; García, Y.C.; Núñez, M.; Díaz, K.; Olea, A.F.; Espinoza, L. Plant Growth-Defense Trade-Offs: Molecular Processes Leading to Physiological Changes. Int. J. Mol. Sci. 2021, 22, 693. [Google Scholar] [CrossRef]
  113. Tsuda, K.; Somssich, I.E. Transcriptional networks in plant immunity. New Phytol. 2015, 206, 932–947. [Google Scholar] [CrossRef]
  114. Huerta-Espino, J.; Singh, R.; Crespo-Herrera, L.A.; Villaseñor-Mir, H.E.; Rodriguez-Garcia, M.F.; Dreisigacker, S.; Barcenas-Santana, D.; Lagudah, E. Adult Plant Slow Rusting Genes Confer High Levels of Resistance to Rusts in Bread Wheat Cultivars from Mexico. Front. Plant Sci. 2020, 11, 824. [Google Scholar] [CrossRef] [PubMed]
  115. Krattinger, S.G.; Lagudah, E.S.; Spielmeyer, W.; Singh, R.P.; Huerta-espino, J.; Mcfadden, H.; Bossolini, E.; Selter, L.L.; Keller, B. Pathogens in Wheat. Science 2009, 323, 1360–1363. [Google Scholar] [CrossRef] [Green Version]
  116. Ellis, J.G.; Lagudah, E.S.; Spielmeyer, W.; Dodds, P.N. The past, present and future of breeding rust resistant wheat. Front. Plant Sci. 2014, 5, 641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Böni, R.H. Functional Characterization of the Wheat Disease Resistance Gene Lr34 in Functional Characterization of the Wheat Disease Resistance Gene Lr34 in Heterologous Barley. Ph.D. Thesis, University of Zurich, Zürich, Switzerland, 2017. [Google Scholar]
  118. Mandalà, G.; Tundo, S.; Francesconi, S.; Gevi, F.; Zolla, L.; Ceoloni, C.; D’Ovidio, R. Deoxynivalenol Detoxification in Transgenic Wheat Confers Resistance to Fusarium Head Blight and Crown Rot Diseases. Mol. Plant-Microbe Interact. 2019, 32, 583–592. [Google Scholar] [CrossRef] [PubMed]
  119. Tundo, S.; Kalunke, R.; Janni, M.; Volpi, C.; Lionetti, V.; Bellincampi, D.; Favaron, F.; D’Ovidio, R. Pyramiding PvPGIP2 and TAXI-III but Not PvPGIP2 and PMEI Enhances Resistance Against Fusarium graminearum. Mol. Plant-Microbe Interact. 2016, 29, 629–639. [Google Scholar] [CrossRef] [Green Version]
  120. Tundo, S.; Janni, M.; Moscetti, I.; Mandalà, G.; Savatin, D.; Blechl, A.; Favaron, F.; D’Ovidio, R. PvPGIP2 Accumulation in Specific Floral Tissues But Not in the Endosperm Limits Fusarium graminearum Infection in Wheat. Mol. Plant-Microbe Interact. 2016, 29, 815–821. [Google Scholar] [CrossRef]
  121. Obbard, D.J.; Gordon, K.H.J.; Buck, A.; Jiggins, F.M. The evolution of RNAi as a defence against viruses and transposable elements. Philos. Trans. R. Soc. B Biol. Sci. 2008, 364, 99–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. 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]
  123. Sidahmed, A.M.E.; Wilkie, B. Endogenous Antiviral Mechanisms of RNA Interference: A Comparative Biology Perspective. Adv. Struct. Saf. Stud. 2010, 623, 3–19. [Google Scholar] [CrossRef]
  124. Gaffar, F.Y.; Koch, A. Catch Me If You Can! RNA Silencing-Based Improvement of Antiviral Plant Immunity. Viruses 2019, 11, 673. [Google Scholar] [CrossRef] [Green Version]
  125. Fahim, M.; Millar, A.; Wood, C.C.; Larkin, P.J. Resistance to Wheat streak mosaic virus generated by expression of an artificial polycistronic microRNA in wheat. Plant Biotechnol. J. 2011, 10, 150–163. [Google Scholar] [CrossRef] [Green Version]
  126. Kis, A.; Tholt, G.; Ivanics, M.; Várallyay, É.; Jenes, B.; Havelda, Z. Polycistronic artificial miRNA-mediated resistance toWheat dwarf virusin barley is highly efficient at low temperature. Mol. Plant Pathol. 2015, 17, 427–437. [Google Scholar] [CrossRef] [PubMed]
  127. Scorza, R.; Callahan, A.; Dardick, C.; Ravelonandro, M.; Polak, J.; Malinowski, T.; Zagrai, I.; Cambra, M.; Kamenova, I. Genetic engineering of Plum pox virus resistance: ‘HoneySweet’ plum—From concept to product. Plant Cell Tissue Organ Cult. (PCTOC) 2013, 115, 1–12. [Google Scholar] [CrossRef]
  128. Koch, A.; Kumar, N.; Weber, L.; Keller, H.; Imani, J.; Kogel, K.-H. Host-induced gene silencing of cytochrome P450 lanosterol C14 -demethylase-encoding genes confers strong resistance to Fusarium species. Proc. Natl. Acad. Sci. USA 2013, 110, 19324–19329. [Google Scholar] [CrossRef] [Green Version]
  129. Cheng, W.; Song, X.-S.; Xiao-Li, Q.; 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] [PubMed]
  130. Panwar, V.; McCallum, B.; Bakkeren, G. Endogenous silencing of P uccinia 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]
  131. Panwar, V.; McCallum, B.; Bakkeren, G. Host-induced gene silencing of wheat leaf rust fungus Puccinia triticina pathogenicity genes mediated by the Barley stripe mosaic virus. Plant Mol. Biol. 2013, 81, 595–608. [Google Scholar] [CrossRef]
  132. Chen, W.; Kastner, C.; Nowara, D.; Oliveira-Garcia, E.; Rutten, T.; Zhao, Y.; Deising, H.B.; Kumlehn, J.; Schweizer, P. Host-induced silencing of Fusarium culmorum genes protects wheat from infection. J Exp Bot. 2016, 67, 4979–4991. [Google Scholar] [CrossRef] [Green Version]
  133. Nelson, R.; Wiesner-Hanks, T.; Wisser, R.; Balint-Kurti, P. Navigating complexity to breed disease-resistant crops. Nat. Rev. Genet. 2018, 19, 21–33. [Google Scholar] [CrossRef] [PubMed]
  134. Mohanta, T.K.; Bashir, T.; Hashem, A.; Allah, E.F.A.; Bae, H. Genome Editing Tools in Plants. Genes 2017, 8, 399. [Google Scholar] [CrossRef] [Green Version]
  135. Arora, L.; Narula, A. Gene Editing and Crop Improvement Using CRISPR-Cas9 System. Front. Plant Sci. 2017, 8, 1932. [Google Scholar] [CrossRef] [Green Version]
  136. Wang, Y.; Cheng, X.; Shan, Q.; Zhang, Y.; Liu, J.; Gao, C.; Qiu, J.-L. Simultaneous editing of three homoeoalleles in hexaploid bread wheat confers heritable resistance to powdery mildew. Nat. Biotechnol. 2014, 32, 947–951. [Google Scholar] [CrossRef]
  137. Zhou, J.; Peng, Z.; Long, J.; Sosso, D.; Liu, B.; Eom, J.-S.; Huang, S.; Liu, S.; Cruz, C.V.; Frommer, W.; et al. Gene targeting by the TAL effector PthXo2 reveals cryptic resistance gene for bacterial blight of rice. Plant J. 2015, 82, 632–643. [Google Scholar] [CrossRef]
  138. Wang, F.; Wang, C.; Liu, P.; Lei, C.; Hao, W.; Gao, Y.; Liu, Y.-G.; Zhao, K. Enhanced Rice Blast Resistance by CRISPR/Cas9-Targeted Mutagenesis of the ERF Transcription Factor Gene OsERF922. PLoS ONE 2016, 11, e0154027. [Google Scholar] [CrossRef]
  139. Ma, J.; Chen, J.; Wang, M.; Ren, Y.; Wang, S.; Lei, C.; Cheng, Z. Sodmergen Disruption of OsSEC3A increases the content of salicylic acid and induces plant defense responses in rice. J. Exp. Bot. 2018, 69, 1051–1064. [Google Scholar] [CrossRef] [Green Version]
  140. 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]
  141. Piffanelli, P.; Ramsay, L.; Waugh, R.; Benabdelmouna, A.; D’Hont, A.; Hollricher, K.; Jørgensen, J.H.; Schulze-Lefert, P.; Panstruga, R. A barley cultivation-associated polymorphism conveys resistance to powdery mildew. Nat. Cell Biol. 2004, 430, 887–891. [Google Scholar] [CrossRef] [Green Version]
  142. Consonni, C.; Humphry, M.E.; Hartmann, H.A.; Livaja, M.; Durner, J.; Westphal, L.; Vogel, J.; Lipka, V.; Kemmerling, B.; Schulze-Lefert, P.; et al. Conserved requirement for a plant host cell protein in powdery mildew pathogenesis. Nat. Genet. 2006, 38, 716–720. [Google Scholar] [CrossRef]
  143. Bai, Y.; Pavan, S.; Zheng, Z.; Zappel, N.F.; Reinstädler, A.; Lotti, C.; De Giovanni, C.; Ricciardi, L.; Lindhout, P.; Visser, R.; et al. Naturally Occurring Broad-Spectrum Powdery Mildew Resistance in a Central American Tomato Accession Is Caused by Loss of Mlo Function. Mol. Plant-Microbe Interact. 2008, 21, 30–39. [Google Scholar] [CrossRef] [Green Version]
  144. Sedeek, K.E.M.; Mahas, A.; Mahfouz, M. Plant Genome Engineering for Targeted Improvement of Crop Traits. Front. Plant Sci. 2019, 10, 114. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Agricultural biotechnology timeline. A timeline showing how biotechnology in agriculture has evolved, changing the ability to develop new crops.
Figure 1. Agricultural biotechnology timeline. A timeline showing how biotechnology in agriculture has evolved, changing the ability to develop new crops.
Plants 10 01146 g001
Figure 2. Biotechnological approaches and their possible involvement to enhance cereal resistance to pathogens.
Figure 2. Biotechnological approaches and their possible involvement to enhance cereal resistance to pathogens.
Plants 10 01146 g002
Table 1. Biotechnological interventions to increase disease resistance in cereals.
Table 1. Biotechnological interventions to increase disease resistance in cereals.
Immunity Level of InterventionBiotechnological InterventionGeneSpeciesEnhanced Resistance toReferences
Pathogen sensingInterspecies/interfamily transfer of known PRRsAtEFRWheatPseudomonas syringae pv. oryzae[79]
AtEFRRiceXanthomonas oryzae pv. oryzae-derived elf18[80]
AtEFRRiceAcidovorax avenae subsp. avenae[81]
OsXa21RiceXanthomonas oryzae pv. oryzae[82]
TaRLK1 and TaRLK2WheatBlumeria graminis f. sp. tritici[83]
HvLEMK1Barely, WheatBlumeria graminis f.sp. hordei; Blumeria graminis f. sp. tritici[84]
HvLecRK-VWheatBlumeria graminis f. sp. tritici[85]
Production of chimeric receptor kinases and R genesAtEFR-OsXa21RicePseudomonas syringae pv. tomato; Agrobacterium tumefaciens; Xanthomonas oryzae pv. oryzae[86,87]
OsXa21-OsCEPiPRiceMagnaporthe oryzae[88]
Effector detectionDeletion of effector binding sitesOs11N3/OsSWEET14RiceXanthomonas oryzae pv. oryzae[89]
Addition of effector binding sitesOsXa27RiceXanthomonas oryzae pv. oryzae[90]
Immune signalingAltered expression of signaling componentsAtNPR1RiceBroad-spectrum of pathogens[91]
Altered expression of transcription factorsTaPIMP1WheatBipolaris sorokiniana[92]
OsIPA1/OsSPL14RiceMagnaporthe oryzae[93]
R genesTransfer of APR allelesTaLr34Barely, Rice, Sorghum Maize, Durum wheatMultiple biotrophic pathogens [94,95,96,97,98]
TaLr67BarelyMultiple rusts and powdery mildew[99]
Table 2. Examples of gene expression or editing techniques to increase disease resistance in cereals.
Table 2. Examples of gene expression or editing techniques to increase disease resistance in cereals.
Molecular TechniqueBiotechnological InterventionGeneSpeciesEnhanced Resistance toReferences
RNAiViral gene silencingWheat streak mosaic virus genesWheatWheat streak mosaic virus (WSMV)[125]
Wheat dwarf virus genesBarelyWheat dwarf virus (WDV)[126]
Host-induced gene silencingFgCYP51A, FgCYP51B and FgCYP51CBarelyFusarium graminearum[128]
FgCh3bWheatFusarium graminearum[129]
PtMAPK1, PtCYC1, PtCNBWheatPuccinia triticina, P. graminis and P. striiformis[130,131]
FcGlsWheatFusarium culmorum[132]
CRISPR/Cas9Silencing of host genesTaMlo-A1WheatBlumeria graminis f. sp. tritici[136]
OsSWEET13RiceXanthomonas oryzae pv. oryzae[137]
OsERF922RiceMagnaporthe oryzae[138]
TaEDR1WheatBlumeria graminis f. sp. tritici[43]
OsSEC3ARiceMagnaporthe oryzae[139]
TaLpx-1WheatFusarium graminearum[102]
TaHRCWheatFusarium graminearum[140]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bigini, V.; Camerlengo, F.; Botticella, E.; Sestili, F.; Savatin, D.V. Biotechnological Resources to Increase Disease-Resistance by Improving Plant Immunity: A Sustainable Approach to Save Cereal Crop Production. Plants 2021, 10, 1146. https://doi.org/10.3390/plants10061146

AMA Style

Bigini V, Camerlengo F, Botticella E, Sestili F, Savatin DV. Biotechnological Resources to Increase Disease-Resistance by Improving Plant Immunity: A Sustainable Approach to Save Cereal Crop Production. Plants. 2021; 10(6):1146. https://doi.org/10.3390/plants10061146

Chicago/Turabian Style

Bigini, Valentina, Francesco Camerlengo, Ermelinda Botticella, Francesco Sestili, and Daniel V. Savatin. 2021. "Biotechnological Resources to Increase Disease-Resistance by Improving Plant Immunity: A Sustainable Approach to Save Cereal Crop Production" Plants 10, no. 6: 1146. https://doi.org/10.3390/plants10061146

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