Next Article in Journal
Eukaryotic CRFK Cells Motion Characterized with Atomic Force Microscopy
Next Article in Special Issue
Overexpression of the Purple Perilla (Perilla frutescens (L.)) FAD3a Gene Enhances Salt Tolerance in Soybean
Previous Article in Journal
Evaluation of Reference Genes for Real-Time Quantitative PCR Analysis in Tissues from Bumble Bees (Bombus Terrestris) of Different Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transgenic Improvement for Biotic Resistance of Crops

Maize Research Institute, Sichuan Agricultural University, Ya’an 611130, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 14370; https://doi.org/10.3390/ijms232214370
Submission received: 25 October 2022 / Revised: 15 November 2022 / Accepted: 17 November 2022 / Published: 19 November 2022

Abstract

:
Biotic constraints, including pathogenic fungi, viruses and bacteria, herbivory insects, as well as parasitic nematodes, cause significant yield loss and quality deterioration of crops. The effect of conventional management of these biotic constraints is limited. The advances in transgenic technologies provide a direct and directional approach to improve crops for biotic resistance. More than a hundred transgenic events and hundreds of cultivars resistant to herbivory insects, pathogenic viruses, and fungi have been developed by the heterologous expression of exogenous genes and RNAi, authorized for cultivation and market, and resulted in a significant reduction in yield loss and quality deterioration. However, the exploration of transgenic improvement for resistance to bacteria and nematodes by overexpression of endogenous genes and RNAi remains at the testing stage. Recent advances in RNAi and CRISPR/Cas technologies open up possibilities to improve the resistance of crops to pathogenic bacteria and plant parasitic nematodes, as well as other biotic constraints.

1. Transgenic Improvement Is Effective to Control Biotic Constraints

1.1. Biotic Constraints to Crops

As sessile organisms, crop plants have to cope with other living organisms in their environment. Except for a few, such as rhizobia, which are beneficial to their hosts, most of the others cause some damage to crop plants such as pathogenic fungi, viruses, bacteria, herbivory insects, as well as parasitic nematodes. They are classified as biotic constraints to crop plants [1,2,3,4,5]. Pathogenic fungi and bacteria are mostly intracellular pathogens. They invade different organs of the plant during the interaction with the host, live inside and kill the host cells using toxic enzymes or compounds such as necrotrophs or initially feed on plants parasitically as biotrophs or hemi-biotrophs, leading to vascular wilts, leaf spots, and cankers, among other symptoms [4,5]. Pathogenic viruses produce not only local lesions but also systemic damage that causes stunting, chlorosis, and malformations, although they rarely kill their hosts. They have a very efficient system of dissemination through vector transmission by insects, arthropods, and nematodes [6]. Herbivory insects, as well as mites, damage plants through feeding or egg-laying. Piercing–sucking insects can also act as virus vectors, transmitting them to plants through their styles [1,7]. Plant-parasitic nematodes attack the root system, withdraw the nutrients of plant cells, and alter plant development, physiology, and immunity, resulting in leaf necrosis and chlorosis, plant wilting, and stunting, and enhanced susceptibility to other pathogens [8,9,10]. Several economically important genera parasitize various crop plants. The root-knot, root lesion, and cyst nematodes of Heteroderidae are the three most economically damaging genera of plant-parasitic nematodes and lead to significant yield loss [10]. In order to deal with these biotic stresses, plants have evolved an advanced immune system, such as physical barriers (waxes, thick cuticles, and specialized trichomes) to prevent pathogen settling and herbivory feeding, chemical compounds to protect against herbivory and pathogen infection, and immune responses that activate plant defense mechanisms in a much more effective way [11,12,13,14,15]. However, these mechanisms cannot completely resist pathogen and herbivory damage, which still affect crop development and lead to a yield loss of 17–30% (Figure 1) and significant quality deterioration, and are irrefutably the most serious threat to worldwide food security [1,7,16,17,18,19]. In the Brassica crops, which can be invaded and fed by numerous pathogens and herbivory insects, the yield loss is estimated as high as 50–60% [20,21,22].

1.2. Conventional Management

To meet the food demand of the booming world population, sustainable agriculture makes great efforts to reduce the yield loss and quality deterioration of crops without compromising the biotic constraints [1,23]. In agricultural practice, the conventional approaches are traditional breeding for resistance and application of agrochemicals (fungicides, bactericides, nematicides, and insecticides), assisted by comprehensive field management and biological control [24]. The application of resistant cultivars with broad-spectrum and durable resistance is the most cost-effective, environmentally friendly, and less labor-intensive [25,26,27,28]. However, traditional breeding heavily depends on germplasm resources. It is effective for the improvement of vertical resistance to specific parasitic pathotypes of biotrophic pathogens, such as rust fungi Puccinia striiformis, P. triticina, and P. graminis (pathogens of stripe, leaf, and stem rust of wheat, respectively) and Phytophthora infestans (pathogen of potato late-blight) [29,30,31,32]. This kind of resistance is complete but can be easily evaded due to the continuous accumulating mutations of pathogens, leading to the breakdown of resistant varieties after their commercial utilization [33,34,35,36]. For improvement of horizontal resistance to non-specific parasitic pathotypes of saprophytic pathogens, such as Bhizoctonia solani and Hyaloperonospora parasitica (pathogens of maize sheath blight and rape downy mildew, respectively), as well as pathogenic viruses and phytophagous insects, it is a great challenge to pyramid the multiple dominant resistant genes into a single cultivar for developing broad-spectrum and durable resistance [37,38,39,40,41,42]. This challenge occurs even with the advances in distant hybridization, artificial mutation, double haploid technology, somatic hybrid, and DNA marker-assisted genotyping and selection [38,39,43,44,45,46,47,48,49], due to the limitation of germplasm resources, the genetic background of minor polygenes, and negative correlation with yield and quality characteristics [50,51,52,53]. The application of agrochemicals is effective for the control of herbivory insects, but increases production costs and causes environmental pollution [54,55,56,57]. For the control of pathogenic fungi and bacteria, especially plant parasitic nematodes, conventional management is much more difficult. Comprehensive field management procedures, such as field sanitation, soil sterilization, crop rotation and fallowing, rouging, and the manual removal of infected plants, are only applied as assistive approaches, because of their labor input and limits in disease and insect eradication [16,58,59]. Biological control may not be compatible with the crop production system or the regulatory environment and remains at the testing stage, except for a few successful stories [60,61].

1.3. Transgenic Improvement

Transgenic technology overcomes the hybridization barriers, utilizes a broader and more diverse range of desirable genes from genetically distant species, and provides a direct and directional approach to improve crops for biotic resistance in a careful target manner with minimal effect on beneficial soil microbes and environment [62,63,64]. It is praised as a second Green Revolution, hopeful for improving biotic resistance and other agronomical characteristics of crops and making a great contribution to food security [65,66]. Transgenic cultivars of crops are generated by the cloning of desirable genes, construction of expression vectors, genetic transformation of recipient cultivars, and identification of transgenic lines, so as to improve the original undesirable traits or endow them with new beneficial traits [65,66,67]. Transgenic technology is also used to modify or knock out some suboptimal genes to change their undesirable phenotypes [68,69,70,71]. After a strict assessment for food and environmental safety, transgenic cultivars with significant improvement in biotic resistance and other agronomical characteristics are authorized for cultivation and market [72]. According to the survey carried out by the International Service for the Acquisition of Agri-Biotech Applications (ISAAA, http://www.isaaa.org/gmapprovaldatabase/default.asp, accessed on 22 September 2022), 292 transgenic events and hundreds of varieties and hybrids between them of 32 cereals, vegetables, fruits, fibers, oils, trees, and forage crops have been authorized for cultivation and market, cultivated for 176.85 million hectares in more than thirty countries, and providing great profitability by increasing yield and reducing input in pesticides, labor, and machinery. They are harmless to humans, environmentally friendly, and favored by farmers worldwide [62,73,74]. A meta-analysis shows that the application of herbicide-resistant and insecticidal transgenic crops reduces the use of synthetic herbicide and pesticides by 2.43% and 41.67%, and their cost by 25.29% and 43.43%, increases the yield of crops by 9.29% and 24.85%, and benefits farmers by 64.29% and 68.78%, respectively (Figure 2) [75].

2. Commercial Release for Transgenic Resistance to Herbivory Insects, and Pathogenic Viruses, and Fungi

2.1. Resistance to Herbivory Insects

One of the most successful transgenic improvements for biotic resistance is the wide application of transgenic events resistant to herbivory insects, which used to be the major biotic constraint that caused a serious reduction in crop productivity and posed a great threat to food security in the world [7]. According to the survey carried out by ISAAA (http://www.isaaa.org/gmapprovaldatabase/default.asp, accessed on 22 September 2022), 93 single transgenic events (not included hybrids between them. The same below.) resistant to herbivory insects of lepidopteran, coleopteran, as well as hemipteran orders, have been authorized for cultivation and market. From them, hundreds of cotton, cowpea, eggplant, maize, potato, rice, soybean, sugarcane, and tomato varieties and hybrids have been developed and authorized for cultivation and market. Their application has reduced pesticide usage and increased yield and the benefits of crop production [73,74,75,76,77,78,79]. Ninety of these ninety-three transgenic events were transformed by the heterologous expression of one or more (for pyramiding resistance) of the insecticidal genes Cry (δ–endotoxin) from different strains of soil bacterium Bacillus thuringiensis [80], and one was transformed by the heterologous expression of the vegetative insecticidal protein gene vip3 also from B. thuringiensis. Ten were simultaneously transformed by orthologs of the Cry gene from the other strains of B. thuringiensis or for pyramiding resistance. Three were simultaneously transformed by vegetative insecticidal protein genes vip3, CpTI, and API, as well as double-stranded RNA transcript of gene Snf7 from Western corn rootworm (Diabrotica virgifera), respectively, for pyramiding resistance (Table 1) [81,82,83,84].

2.2. Resistance to Pathogenic Viruses

In the early years, the transgenic improvement for resistance to pathogenic viruses was focused on pathogen-derived resistance (PDR), whereby a sequence or part of the viral genome was introduced into the host plant [85,86,87,88,89]. After its discovery in the model non-parasitic nematode Caenorhabditis elegans and demonstration in transgenic improvement of vial resistance [90,91], RNA interference (RNAi) technology has rapidly emerged as the most effective strategy of transgenic improvement for resistance to plant viruses [91]. Synthetic DNA sequences expressing double-stranded RNAs (dsRNAs) for interference of the housekeeping genes of pathogenic viruses, such as the genes of the coat protein (CP) [92,93,94], the replication protein (Rep) [95,96,97], as well as the movement protein (MP) [98,99]. According to the survey carried out by ISAAA (http://www.isaaa.org/gmapprovaldatabase/default.asp, accessed on 22 September 2022), twenty-five transgenic events of bean, papaya, plum, potato, squash, sweet pepper, and tomato resistant to pathogenic viruses were generated and authorized for cultivation and market (Table 2). All of them were transformed by double-stranded sequences transcribing the coat protein, replicase, or helicase of the pathogenic viruses themselves for RNAi [94,100,101,102]. In recent years, clustered regularly interspaced short palindromic repeats-cas (CRISPR/Cas) technology and its new advances attracted much attention in transgenic improvement for rival resistance [103,104,105,106]. Transgenic potato lines with broad-spectrum resistance were developed by introduction of a multiple gRNA cassette of CRSIPR/Cas13 system to target six genes of three potato virus strains [106].

2.3. Resistance to Pathogenic Fungi

A lot of attempts have been made to battle against a wide range of pathogenic fungi by transgenic improvement [107,108,109,110,111,112,113,114,115]. However, most these attempts were generated by the overexpression of endogenous genes resistant to pathogenic fungi [109,116,117]. Unfortunately, their resistance improvement was usually non-significant, highly specific against a few pathogens, and even showed reduced yield or growth vigor, due to the complex mechanisms and signaling networks of disease resistance [107,109,118]. Almost all of them remain at testing stage and few commercial applications have yet been achieved [107,119]. Much effort has been also paid to the heterologous expression of exogenous resistant genes. The immune receptor gene AtRLP23 was introduced into potato and resulted in smaller lesions and reduced pathogen colonization when infected with Phytophthora infestans or Sclerotinia sclerotiorum [120]. The heterologous expression of the immune receptor genes SlVe1 from tomato and SmELR from wild potato also enhanced resistance to Verticillium wilt in tobacco and cotton, and Phytophthora infestans in potato, respectively [121,122]. According to the survey carried out by ISAAA (http://www.isaaa.org/gmapprovaldatabase/default.asp, accessed on 22 September 2022), only four transgenic events for the resistance improvement for pathogenic fungi have been authorized for cultivation and market for their significant improvement of resistance to the potato late-blight pathogen (Phytophthora infestans). One (event SP951) was transformed by the heterologous expression of exogenous resistant genes RB from distant species Solanum bulbocastanum of the nightshade family. The resistance of the other three (W8, X17, and Y9) were conferred by the heterologous expression of exogenous gene Rpi-vnt from distant species Solanum venturi of the nightshade family, and pyramided by RNAi of endogenous genes Asn1, Ppo5, Phl, R1, and Vinv related to pathogenesis [123]. In recent times, a more efficient strategy was developed to combine multiple resistance genes in an expression cassette. These combined genes not only increased the resistance spectrum to fungal rust pathogen in the transgenic wheat lines, but also inherited a single genetic locus without segregation [124].

3. Exploration for Transgenic Resistance to Bacteria and Nematodes

3.1. Resistance to Pathogenic Bacteria

A lot of efforts have also been devoted to the creation of transgenic resistance to bacterial diseases [107,125,126]. In fact, antibacterial proteins, widely present in animals, plants, and bacteriophages, have a bactericidal action on a large range of Gram-negative and -positive bacteria. Their encoding genes have been cloned from genetic distant species and heterogeneously expressed in crops in attempts to confer resistance to bacterial diseases [125]. For example, the genes of cecropins (lytic peptides) were cloned from the giant silk moth and expressed in potato and tobacco plants. The transgenic lines showed delayed symptoms and reduced mortality following inoculation with Ralstonia solanacearum and Pseudomonas solanacearum, and Pseudomonas syringae (pathogen of bacterial wilt), respectively [127,128]. The heterogenous expression of the genes of lysozymes, lactoferrin, chitinase, the effectors of salicylic acid and jasmonic acid signaling, as well as other antibacterial peptide or phytoalexins genes, and silencing the expression of toxins, pectic enzymes, and exopolysaccharides genes of pathogenic bacteria have also been tried in the transgenic improvement of crops for bacterial resistance [129,130,131,132,133,134,135]. For example, the heterologous expression of the EF-Tu receptor gene AtEFR of Arabidopsis in tomato, potato, tobacco, wheat, and rice enhanced resistance to different pathogenic bacteria [136,137,138,139,140,141,142]. However, the safety assessment for the efficacy, durability, absence of toxicity, and low environmental impact of these transgenic events are much more difficult than other transgenes. None of them have been authorized for cultivation and market [125].

3.2. Resistance to Parasitic Nematodes

Although overexpression of endogenous and exogenous resistance genes, such as the Mi from tomato for resistance against Meloidogyne incognita, the pro-1 from sugar beet against Heterodera schachtii, the Gpa-2 from potato against Globodera allida, and the Hero from tomato against Globodera rostochiensis, as well as the genes of cowpea trypsin inhibitor, cystatins, and serine proteases, showed some enhancement to the resistance of the transgenic lines [143]; the most effective strategy of transgenic improvement for resistance to parasitic nematodes is RNAi, which was first discovered in the model non-parasitic nematode C. elegans and also demonstrated in plant-parasitic nematodes [144,145,146]. Crop plants are transformed by RNAi constructs to express dsRNA molecules with sequences derived from the target genes related to parasitism and the development of nematodes, such as dual oxidase, splicing factor, integrase, cathepsin L cysteine proteinase, proteasome integrity-related protein, signal peptidase, tyrosine phosphatase, neuropeptide, as well as aspartic, serine, cysteine proteases, and effector in M. incognita [147,148,149,150,151,152,153,154,155,156], Pratylenchus vulnus [157], and Meloidogyne chitwoodi [158], and protease and cyst formation- and reproduction-associated proteins, as well as three genes associated with reproduction to fitness in Heterodera glycines [159,160]. Among them, the effector genes should be a worthwhile group of target genes for in planta RNAi strategy as effectors generally lack high homology with the genes of organisms from other taxa, thereby diminishing the potential for problems related to off-target effects [161].

4. Strategy Option of Transgenic Improvement

4.1. Three Strategies of Transgenic Manipulation

The heterologous expression of exogenous genes from genetically distant species, overexpression of endogenous genes from the same species or homologous genes from sexually compatible species, and silencing of the expression of suboptimal endogenous, pathogenic, or pest genes by RNAi or CRISPR/Cas are three strategies of transgenic improvement. The first strategy is the transformation by exogenous genes from genetically distant species. The second strategy is the transformation of endogenous genes from the same species or homologous genes from sexually compatible species. The third strategy is silencing the expression of undesirable endogenous or pathogenic and pest genes by RNAi) or CRISPR/Cas technology [162].

4.2. Heterologous Expression of Exogenous Genes

The exogenous transgenes were promoted by the constitutive promoters and the possible codon usage bias was usually overcome by codon optimization of the transgene sequences [163,164]. Therefore, their heterologous expression was usually not regulated on the transcriptional level, although the expression levels might be affected by genetic background and growth stage of the recipient cultivars, as well as environmental conditions [77,165,166,167]. The investigations in transgenic insecticidal cotton (r = 0.762, p < 0.001) and rice (r = 0.742, p < 0.01) show that the accumulation of the Cry protein in leaves is nonlinearly correlated with the heterologous transcription levels of the exogenous Cry genes, although varying with growth and development [168,169,170,171]. Therefore, the novel functions conferred by the exogenous transgenes, such as resistance to herbivory insects [74,76,77,78,79,80], usually performed in pathways divergent from the endogenous metabolism of the recipient crops [172,173,174]. This strategy should be hopeful for transgenic improvement for broad-spectrum resistance to pathogenic fungi with the detailed elucidation of the mechanism of the hypersensitive response mediated by the non-host resistant or avirulence genes [175,176,177,178].

4.3. Overexpression of Endogenous Genes

Biochemical reactions are reversible and regulated in complex networks [80,179]. The activity of any step enhanced by overexpression of the gene encoding the enzyme may not desirably result in increased flux through the reaction that it catalyzes [77,180]. Moreover, transformation may generate completely new interactions between the transgenes, making them function differently from what is expected. According to the survey carried out by ISAAA (http://www.isaaa.org/gmapprovaldatabase/default.asp, accessed on 22 September 2022), 93 of the 118 singular transgenic events of biotic resistance authorized for cultivation and market were created by the first strategy and transformed by bacterial insect-resistant genes [77,179,180,181], except one exogenous gene from a sexually incompatible plant species was introduced into one event for pyramiding herbicide resistance [84]. The other 25 events were created by the third strategy and transformed by synthetic sequences transcribing antisense or double-stranded RNAs for silencing the expression of housekeeping genes of pathogenic viruses [92,93,94,97,100,101,102,182,183,184,185]. Of course, antibiotic or herbicide-resistant genes from bacteria were also introduced into almost all of these events as selection markers of transformant screening [186]. Although a lot of the literature has documented the improvement of biotic resistance by overexpression of endogenous genes or homologous genes [187,188,189,190,191,192], all of them remain at the testing stage and none of them have been released commercially through safety evaluation. It is questionable whether this strategy will ever be widely used in transgenic improvement for biotic constraints [107].

4.4. RNA Interference

RNAi triggered by antisense or double-stranded RNAs described by transformed synthetic DNA sequences is a versatile, effective, safe, and eco-friendly technology for crop protection against viruses and other pathogens as well as herbivory insect and nematodes [71,193,194,195]. Because of the advantages, as it does not produce any functional foreign proteins and targets organisms in a sequence-specific manner, host-delivered RNAi has become a powerful tool of transgenic improvement for biotic constraints [71,193,194,195,196]. Up to now, 25 transgenic events of resistant to pathogenic viruses have been authorized for cultivation and market (Table 2) [93,100,101,102]. In particular, the transgenic papaya cultivars resistant to ringspot virus, that is a devastating disease, have rescued the entire production of papaya [182]. The creation of a topical application of dsRNA using layered double-hydroxide clay nanosheets opens up possibilities to exploit such innovations for specific and combinatorial resistance against herbivory insects and plant parasitic nematodes [197]. However, off-target effects can originate due to sequence similarity between dsRNA and non-target mRNA transcripts resulting in an unintended silence of non-target genes and an unexpected phenotype [198,199,200]. This is an important consideration for biosafety assessment of transgenic events of host-generating RNAi. Therefore, proper designing of the dsRNA sequence to preventing the off-target effects, as well as probable non-target effects, is crucial for the wider employment of RNAi. The advancement of genome sequence data, functional genomics availability, and new bioinformatic tools have become helpful for the precise design of dsRNA sequence [201]. RNAi strategy is also hopeful for transgenic improvement for resistance to pathogenic fungi and bacteria with the detailed elucidation of the crucial factors of pathogenicity [202,203,204].

4.5. Gene Edition

After zinc finger nuclease (ZFN) and transcription activator-like effector nuclease (TALEN), CRISPR/Cas technology emerged as a precise, simple, and cost-effective tool of RNA-guided genome editing and is now becoming widely applied to edit the genomes of a diverse range of crop plants [205,206,207,208]. Without incorporating any foreign DNA, it is even proposed as non-transgenic genetic modification fostering public acceptance [209,210]. At the beginning, most studies focused more on the concept proofing of the CRISPR/Cas system [211,212,213,214,215,216]. In recent years, the probable off-target effects caused by the imperfect matches with gRNA and the unpredictable efficiency among different DNA target sites and PAM were effectively reduced with the advances in high attractive sgRNA, high fidelity Cas 9, and transformant screening [206,217]. This system has also been applied to improve crops for biotic resistance, such as overexpressing hypersensitivity-responsive genes of host crops, suppressing effector genes of pathogens and susceptibility genes of host crops, such as virus resistance [112,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233]. For example, bacterial blight disease, caused by Xanthomonas oryzae pv. oryzae, is a major disease that affects rice production. The expression of sucrose-transporter genes SWEET1, SWEET3, and SWEET14 causes disease susceptibility. CRISPR/Cas-editing inhibited the activation of the susceptibility gene (S) of X. oryzae and prevented its encoding for transcription activator-like effectors (TALEs). The SWEET genes could not be activated by the binding of the TALES to their effector binding elements (EBEs) during their promoters and prevented the establishment of host susceptibility [234,235]. In recent years, more and more biotic resistance has been included in the ranks of CRISPR/Cas-mediated improvements (Table 3) [104,208,234,235,236,237,238,239,240,241,242]. However, only a very few events have been in the pipeline of safety assessment up to now [222,224]. Several events have successfully skipped the regulation of government and entered the market because of their safety assurance, and some more events have been in the pipeline of safety assessment [227,228,243]. Recently, CRISPR/Cas13 was characterized as an RNA-guided RNA editing system with high cleavage activities to foreign RNAs [244]. This system can precisely target plant RNA viruses to impart resistance, and even can be multiplexed by designing a streamlined multiple gRNA cassette [105,245]. By this strategy, three transgenic potato lines with broad-spectrum resistance were developed by the introduction of a multiple gRNA cassette of CRSIPR/Cas13 system to target six genes of three potato virus strains [106].

5. Conclusions

Transgenic manipulation is a direct and directional approach to improve crops for biotic resistance. The heterologous expression of exogenous genes from genetically distant species, as well as silencing the expression by RNAi and CRISPR/Cas is more effective than overexpression of endogenous genes. Recent advances in RNAi and CRISPR/Cas technologies open up possibilities to improve the resistance of crops to pathogenic bacteria and plant parasitic nematodes, as well as the other biotic constraints, and enable the preclusion of public concerns and engendering of public acceptance.

Author Contributions

Original draft preparation, H.Y.; reference search and data curation, Y.W.; supervision and funding acquisition, F.F. and W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Department of Science and Technology of Sichuan Province (2022YFH0067), the National Natural Science Foundation of China (32102226), and the Science and Technology Bureau of Chengdu (2021-YF05-02024-SN).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We are grateful to the Key Laboratory of Biology and Genetic Improvement of Maize in Southwest Region, Ministry of Agriculture, for its technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oerke, E.C. Crop losses to pests. J. Agric. Sci. 2006, 144, 31–43. [Google Scholar] [CrossRef]
  2. Gimenez, E.; Salinas, M.; Manzano-Agugliaro, F. Worldwide research on plant defense against biotic stresses as improvement for sustainable agriculture. Sustainability 2018, 10, 391. [Google Scholar] [CrossRef] [Green Version]
  3. McKenna, T.P.; Koziol, L.; Bever, J.D.; Crews, T.E.; Sikes, B.A. Abiotic and biotic context dependency of perennial crop yield. PLoS ONE 2020, 15, e0234546. [Google Scholar] [CrossRef]
  4. Nazarov, P.A.; Baleev, D.N.; Ivanova, M.I.; Sokolova, L.M.; Karakozova, M.V. Infectious plant diseases: Etiology, current status, problems and prospects in plant protection. Acta Nat. 2020, 12, 46–59. [Google Scholar] [CrossRef] [PubMed]
  5. Schumann, G.L.; D’Arcy, C.J. Essential Plant Pathology; APS Press: St. Paul, MN, USA, 2006; pp. 1–338. [Google Scholar]
  6. Falk, B.W.; Nouri, S. Special issue: Plant virus pathogenesis and disease control. Viruses 2020, 12, 1049. [Google Scholar] [CrossRef]
  7. Douglas, A.E. Strategies for enhanced crop resistance to insect pests. Annu. Rev. Plant Biol. 2018, 69, 637–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Jones, J.T.; Haegeman, A.; Danchin, E.G.J.; Gaur, H.S.; Helder, J.; Jones, M.G.K.; Kikuchi, T.; Manzanilla-López, R.; Palomares-Rius, J.E.; Wesemael, W.M.; et al. Top 10 plant-parasitic nematodes in molecular plant pathology. Mol. Plant Pathol. 2013, 14, 946–961. [Google Scholar] [CrossRef] [Green Version]
  9. Elling, A.A. Major emerging problems with minor Meloidogyne species. Phytopathology 2013, 103, 1092–1102. [Google Scholar] [CrossRef] [Green Version]
  10. Ali, M.A.; Azeem, F.; Abbas, A.; Joyia, F.A.; Li, H.; Dababat, A.A. Transgenic strategies for enhancement of nematode resistance in plants. Front. Plant Sci. 2017, 8, 750. [Google Scholar] [CrossRef] [Green Version]
  11. Jones, D.G.; Dangl, L. The plant immune system. Nature 2006, 444, 323–329. [Google Scholar] [CrossRef]
  12. Schreiber, K.; Desveaux, D. Message in a bottle: Chemical biology of induced disease resistance in plants. Plant Pathol. J. 2008, 24, 245–268. [Google Scholar] [CrossRef]
  13. 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] [PubMed]
  14. Monaghan, J.; Zipfel, C. Plant pattern recognition receptor complexes at the plasma membrane. Curr. Opin. Plant Biol. 2012, 15, 349–357. [Google Scholar] [CrossRef] [PubMed]
  15. Muthamilarasan, M.; Prasad, M. Plant innate immunity: An updated insight into defense mechanism. J. Biosci. 2013, 38, 433–449. [Google Scholar] [CrossRef]
  16. Savary, S.; Willocquet, L.; Pethybridge, S.J.; Esker, P.; McRoberts, N.; Nelson, A. The global burden of pathogens and pests on major food crops. Nat. Ecol. Evol. 2019, 3, 430–439. [Google Scholar] [CrossRef]
  17. Savary, S.; Willocquet, L. Modeling the impact of crop diseases on global food security. Ann. Rev. Phytopathol. 2020, 8, 24. [Google Scholar] [CrossRef]
  18. Hampf, A.C.; Nendel, C.; Strey, S.; Strey, R. Biotic yield losses in the Southern Amazon, Brazil: Making use of smartphone–assisted plant disease diagnosis data. Front. Plant Sci. 2021, 12, 621168. [Google Scholar] [CrossRef]
  19. Fontana, D.C.; de Paula, S.; Torres, A.G.; de Souza, V.H.M.; Pascholati, S.F.; Schmidt, D.; Dourado Neto, D. Endophytic fungi: Biological control and induced resistance to phytopathogens and abiotic stresses. Pathogens 2021, 10, 570. [Google Scholar] [CrossRef]
  20. Kim, C.; Cho, W.; Kim, H. Yield loss of spring Chinese cabbage as affected by infection time of clubroot disease in fields. Plant Dis. Res. 2000, 6, 23–26. [Google Scholar]
  21. Shukla, A.K. Estimation of yield losses to Indian mustard (Brassica juncea) due to Sclerotinia stem rot. J. Phytol. Res. 2005, 18, 267–268. [Google Scholar]
  22. Poveda, J.; Francisco, M.; Cartea, M.E.; Velasco, P. Development of transgenic Brassica crops against biotic stresses caused by pathogens and arthropod pests. Plants 2020, 9, 1664. [Google Scholar] [CrossRef]
  23. Barrett, C.B. Overcoming global food security challenges through science and solidarity. Am. J. Agric. Econ. 2021, 103, 422–447. [Google Scholar] [CrossRef]
  24. Roth, M.G.; Webster, R.W.; Mueller, D.S.; Chilvers, M.I.; Faske, T.R.; Mathew, F.M.; Bradley, C.A.; Damicone, J.P.; Kabbage, M.; Smith, D.L. Integrated management of important soybean pathogens of the United States in changing climate. J. Integr. Pest Manag. 2020, 11, 17. [Google Scholar] [CrossRef]
  25. Yin, K.; Qiu, J.L. Genome editing for plant disease resistance: Applications and perspectives. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2019, 374, 20180322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tester, M.; Langridge, P. Breeding technologies to increase crop production in a changing world. Science 2010, 327, 818–822. [Google Scholar] [CrossRef]
  27. 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]
  28. Sharma, A.; Abrahamian, P.; Carvalho, R.; Choudhary, M.; Paret, M.L.; Vallad, G.E.; Jones, J.B. Future of bacterial disease management in crop production. Annu. Rev. Phytopathol. 2022, 60, 259–282. [Google Scholar] [CrossRef] [PubMed]
  29. Li, C. Breeding crops by design for future agriculture. J. Zhejiang Univ. Sci. B 2020, 21, 423–425. [Google Scholar] [CrossRef]
  30. Roelfs, A.P. Genetic control of phenotypes in wheat stem rust. Ann. Rev. Phytopathol. 1988, 26, 351–367. [Google Scholar] [CrossRef]
  31. Elshafei, A.A.; Motawei, M.I.; Esmail, R.M.; Al-Doss, A.A.; Hussien, A.M.; Ibrahim, E.I.; Amer., M.A. Molecular breeding for rust resistance in wheat genotypes. Mol. Biol. Rep. 2021, 48, 731–742. [Google Scholar] [CrossRef]
  32. Denes, T.E.; Molnar, I.; Rakosy-Tican, E. New insights into the interaction between cultivated potato and Phytophthora infestans. Stud. Univ. Babes-Bolyai Biol. 2015, 60, 165–175. [Google Scholar]
  33. Joosten, M.H.A.J.; Cozijnsen, T.J.; de Wit, P.J.G.M. Host resistance to fungal tomato pathogen lost by a single base pair change in an avirulence gene. Nature 1994, 367, 384–386. [Google Scholar] [CrossRef]
  34. Gassmann, W.; Dahlbeck, D.; Chesnokova, O.; Minsavage, G.V.; Jones, J.B.; Staskawicz, B.J. Molecular evolution of virulence in natural field strains of Xanthomonas campestris pv. vesicatoria. J. Bacteriol. 2000, 182, 7053–7059. [Google Scholar] [CrossRef] [PubMed]
  35. Stukenbrock, E.H.; McDonald, B.A. Population genetics of fungal and oomycete effectors involved in gene-for-gene interactions. Mol. Plant Microbe Interact. 2009, 22, 371–380. [Google Scholar] [CrossRef] [Green Version]
  36. Dodds, P.; Thrall, P. Recognition events and host-pathogen co-evolution in gene-for-gene resistance to flax rust. Funct. Plant Biol. 2009, 36, 395–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Solomon-Blackburn, R.M. Progress in breeding potatoes for resistance to virus diseases. Asp. Appl. Biol. 1998, 52, 299–304. [Google Scholar]
  38. Liu, J.; Liu, D.; Tao, W.; Li, W.; Wang, S.; Chen, P.; Cheng, S.; Gao, D. Molecular marker-facilitated pyramiding of different genes for powdery mildew resistance in wheat. Plant Breed. 2000, 119, 21–24. [Google Scholar] [CrossRef]
  39. Richardson, K.L.; Vales, M.I.; Kling, J.G.; Mundt, C.C.; Hayes, P.M. Pyramiding and dissecting disease resistance QTL to barley stripe rust. Theor. Appl. Genet. 2006, 113, 485–495. [Google Scholar] [CrossRef]
  40. Saghai Maroof, M.A.; Jeong, S.C.; Gunduz, I.; Tucker, D.M.; Buss, G.R.; Tolin, S.A. Pyramiding of soybean mosaic virus resistance genes by marker-assisted selection. Crop Sci. 2008, 48, 517–526. [Google Scholar] [CrossRef]
  41. Keane, P.J. Horizontal or Generalized Resistance to Pathogens in Plants. In Plant Pathology; Cumagun, C.J., Ed.; InTech: Rijeka, Croatia, 2012; pp. 327–362. [Google Scholar]
  42. Galvez, L.C.; Banerjee, J.; Pinar, H.; Mitra, A. Engineered plant virus resistance. Plant Sci. 2014, 228, 11–25. [Google Scholar] [CrossRef]
  43. Mendes, B.J.; Filho, F.M.; Farias, P.; Benedito, V. Citrus somatic hybridization with potential for improved blight and CTV resistance. In Vitro Cell. Dev. Biol. Plant 2001, 37, 490–495. [Google Scholar]
  44. Collard, B.C.; Mackill, D.J. Marker-assisted selection: An approach for precision plant breeding in the twenty-first century. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 2008, 363, 557–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Asea, G.; Vivek, B.S.; Bigirwa, G.; Lipps, P.E.; Pratt, R.C. Validation of consensus quantitative trait loci associated with resistance to multiple foliar pathogens of maize. Phytopathology 2009, 99, 540–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Bohar, R.; Chitkineni, A.; Varshney, R.K. Genetic molecular markers to accelerate genetic gains in crops. Biotechniques 2020, 69, 158–160. [Google Scholar] [CrossRef]
  47. Webster, R.W.; Roth, M.G.; Reed, H.; Mueller, B.; Groves, C.L.; McCaghey, M.; Chilvers, M.I.; Mueller, D.S.; Kabbage, M.; Smith, D.L. Identification of soybean (Glycine max) check lines for evaluating genetic resistance to sclerotinia stem rot. Plant Dis. 2021, 105, 2189–2195. [Google Scholar] [CrossRef]
  48. Foria, S.; Magris, G.; Jurman, I.; Schwope, R.; de Candido, M.; de Luca, E.; Ivanisevic, D.; Morgante, M.; Di Gaspero, G. Extent of wild-to-crop interspecific introgression in grapevine (Vitis inifera) as a consequence of resistance breeding and implications for the crop species definition. Hortic. Res. 2022, 9, uhab010. [Google Scholar] [CrossRef]
  49. Purankar, M.V.; Nikam, A.A.; Devarumath, R.M.; Suprasanna, P. Radiation induced mutagenesis, physio-biochemical profiling and field evaluation of mutants in sugarcane Cv. CoM 0265. Int. J. Radiat. Biol. 2022, 98, 1261–1276. [Google Scholar] [CrossRef]
  50. Batsa, B.K.; Sharma, R.C.; Rai, S.N. Identifying resistance to banded leaf and sheath blight of maize. Indian Phytopathol. 2005, 58, 121–122. [Google Scholar]
  51. Kang, B.C.; Yeam, I.; Jahn, M.M. Genetics of plant virus resistance. Annu. Rev. Phytopathol. 2005, 43, 581–621. [Google Scholar] [CrossRef] [Green Version]
  52. Holbein, J.; Grundler, F.M.; Siddique, S. Plant basal resistance to nematodes: An update. J. Exp. Bot. 2016, 67, 2049–2061. [Google Scholar] [CrossRef] [Green Version]
  53. Preston, G.M. Profiling the extended phenotype of plant pathogens: Challenges in bacterial molecular plant pathology. Mol. Plant Pathol. 2017, 18, 443–456. [Google Scholar] [CrossRef] [Green Version]
  54. Aktar, W.; Sengupta, D.; Chowdhur, Y.A. Impact of pesticides use in agriculture: Their benefits and hazards. Interdiscip. Toxicol. 2009, 2, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Damalas, C.A.; Eleftherohorinos, I.G. Pesticide exposure, safety issues, and risk assessment indicators. Int. J. Environ. Res. Public Health 2011, 8, 1402–1419. [Google Scholar] [CrossRef]
  56. Birkett, M.A.; Pickett, J.A. Prospects of genetic engineering for robust insect resistance. Curr. Opin. Plant Biol. 2014, 19, 59–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Sharma, A.; Kumar, V.; Shahzad, B.; Tanveer, M.; Sidhu, G.P.S.; Handa, N.; Kohli, S.K.; Yadav, P.; Bali, A.S.; Parihar, R.D.; et al. Worldwide pesticide usage and its impacts on ecosystem. SN Appl. Sci. 2019, 1, 1446. [Google Scholar] [CrossRef] [Green Version]
  58. Castle, S.; Palumbo, J.; Prabhaker, N. Newer insecticides for plant virus disease management. Virus Res. 2009, 141, 131–139. [Google Scholar] [CrossRef]
  59. Dababat, A.; Imren, M.; Erginbas-Orakci, G.; Ashrafi, S.; Yavuzaslanoglu, E.; Toktay, H.; Pariyar, S.R.; Elekcioglu, H.I.; Morgounov, A.; Mekete, T. The importance and management strategies of cereal cyst nematodes, Heterodera spp., in Turkey. Euphytica 2015, 202, 173–188. [Google Scholar] [CrossRef]
  60. Bardin, M.; Ajouzm, S.; Comby, M.; Lopez-Ferber, M.; Graillot, B.; Siegwart, M.; Nicot, P.C. Is the efficacy of biological control against plant diseases likely to be more durable than that of chemical pesticides? Front. Plant Sci. 2015, 6, 566. [Google Scholar] [CrossRef]
  61. He, D.C.; He, M.H.; Amalin., D.M.; Liu, W.; Alvindia, D.G.; Zhan, J. Biological control of plant diseases: An evolutionary and eco-economic consideration. Pathogens 2021, 10, 1311. [Google Scholar] [CrossRef] [PubMed]
  62. Raymond Park, J.; Mcfarlane, I.; Hartley Phipps, R.; Ceddia, G. The role of transgenic crops in sustainable development. Plant Biotechnol. J. 2011, 9, 2–21. [Google Scholar] [CrossRef]
  63. Kamthan, A.; Chaudhuri, A.; Kamthan, M.; Datta, A. Genetically modified (GM) crops: Milestones and new advances in crop improvement. Theor. App. Genet. 2016, 129, 1639–1655. [Google Scholar] [CrossRef] [PubMed]
  64. Ahmar, S.; Gill, R.A.; Jung, K.H.; Faheem, A.; Qasim, M.U.; Mubeen, M.; Zhou, W. Conventional and molecular techniques from simple breeding to speed breeding in crop plants: Recent advances and future outlook. Int. J. Mol. Sci. 2020, 21, 2590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Eckardt, N.A.; Cominelli, E.; Galbiati, M.; Tonelli, C. The future of science: Food and water for life. Plant Cell 2009, 21, 368–372. [Google Scholar] [CrossRef] [PubMed]
  66. Farre, G.; Ramessar, K.; Twyman, R.M.; Capell, T.; Christou, P. The humanitarian impact of plant biotechnology: Recent breakthroughs vs bottlenecks for adoption. Curr. Opin. Plant Biol. 2010, 13, 219–225. [Google Scholar] [CrossRef]
  67. Anjanappa, R.B.; Gruissem, W. Current progress and challenges in crop genetic transformation. J. Plant Physiol. 2021, 261, 153411. [Google Scholar] [CrossRef] [PubMed]
  68. Georges, F.; Ray, H. Genome editing of crops: A renewed opportunity for food security. GM Crops Food 2017, 8, 1–12. [Google Scholar] [CrossRef]
  69. Zhang, J.; Khan, S.A.; Heckel, D.G.; Bock, R. Next-generation insect-resistant plants: RNAi-mediated crop protection. Trends Biotechnol. 2017, 35, 871–882. [Google Scholar] [CrossRef]
  70. Mamta, B.; Rajam, M.V. RNAi technology: A new platform for crop pest control. Physiol. Mol. Biol. Plants 2017, 23, 487–501. [Google Scholar] [CrossRef]
  71. Hernández-Soto, A.; Chacón-Cerdas, R. RNAi crop protection advances. Int. J. Mol. Sci. 2021, 22, 12148. [Google Scholar] [CrossRef]
  72. Xu, R.; Zheng, Z.; Jiao, G. Safety assessment and detection methods of genetically modified organisms. Recent Pat. Food Nutr. Agric. 2014, 6, 27–32. [Google Scholar] [CrossRef]
  73. Tilgam, J.; Kumar, K.; Jayaswal, D.; Choudhury, S.; Kumar, A.; Jayaswall, K.; Saxena, A.K. Success of microbial genes based transgenic crops: Bt and beyond Bt. Mol. Biol. Rep. 2021, 48, 8111–8122. [Google Scholar] [CrossRef] [PubMed]
  74. Smyth, S.; Phillips, P.W.B.; Castle, D. Benefits of genetically modified herbicide tolerant canola in Western Canada. Int. J. Biotechnol. 2014, 13, 181. [Google Scholar] [CrossRef]
  75. Klumper, W.; Qaim, M. A meta-analysis of the impacts of genetically modified crops. PLoS ONE 2014, 9, e111629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pray, C.; Ma, D.; Huang, J.; Qiao, F. Impact of Bt cotton in China. World Dev. 2001, 29, 813–825. [Google Scholar] [CrossRef]
  77. Showalter, A.M.; Heuberger, S.; Tabashnik, B.E.; Carriere, Y.; Coates, B. A primer for using transgenic insecticidal cotton in developing countries. J. Insect Sci. 2009, 9, 22. [Google Scholar] [CrossRef] [Green Version]
  78. Blanco, C.A. Heliothis virescens and Bt cotton in the United States. GM Crops Food 2012, 3, 201–212. [Google Scholar] [CrossRef]
  79. Rocha-Munive, M.G.; Soberon, M.; Castaneda, S.; Niaves, E.; Scheinvar, E.; Eguiarte, L.E.; Mota-Sánchez, D.; Rosales-Robles, E.; Nava-Camberos, U.; Martinez-Carrillo, J.L.; et al. Evaluation of the impact of genetically modified cotton after 20 years of cultivation in Mexico. Front. Bioeng. Biotechnol. 2018, 6, 82. [Google Scholar] [CrossRef]
  80. Ghareyazie, B.; Alinia, F.; Menguito, C.A.; Rubia, L.G.; de Palma, J.M.; Liwanag, E.A.; Cohen, M.B.; Khush, G.S.; Bennett, J. Enhanced resistance to two stem borers in an aromatic rice containing a synthetic cryIA(b) gene. Mol. Breed. 1997, 3, 401–414. [Google Scholar] [CrossRef]
  81. Xie, Z.W.; Luo, M.J.; Xu, W.F.; Chi, C.W. Two reactive site locations and structure-function study of the arrowhead proteinase inhibitors, A and B, using mutagenesis. Biochemistry 1997, 36, 5846–5852. [Google Scholar] [CrossRef]
  82. Hu, J.J.; Tian, Y.C.; Han, Y.F.; Li, Y.; Zhang, B.E. Field evaluation of insect-resistant transgenic Populus nigra trees. Euphytica 2001, 121, 123–127. [Google Scholar] [CrossRef]
  83. Cui, J.; Luo, J.; van der Werf, W.; Ma, Y.; Xia, J. Effect of pyramiding Bt and CpTI genes on resistance of cotton to Helicoverpa armigera (Lepidoptera: Noctuidae) under laboratory and field conditions. J. Econ. Entomol. 2011, 104, 673–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ramaseshadri, P.; Segers, G.; Flannagan, R.; Wiggins, E.; Clinton, W.; Ilagan, O.; McNulty, B.; Clark, T.; Bolognesi, R. Physiological and cellular responses caused by RNAi-mediated suppression of Snf7 orthologue in western corn rootworm (Diabrotica virgifera virgifera) larvae. PLoS ONE 2013, 8, e54270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Sanford, J.C.; Johnston, S.A. The concept of parasite-derived resistance- deriving resistance genes from the parasite’s own genome. J. Theor. Biol. 1985, 113, 395–405. [Google Scholar] [CrossRef]
  86. Abel, P.P.; Nelson, R.S.; De, B.; Hoffmann, N.; Rogers, S.G.; Fraley, R.T.; Beachy, R.N. Delay of disease development in transgenic plants that express the tobacco mosaic virus coat protein gene. Science 1986, 232, 738–743. [Google Scholar] [CrossRef] [PubMed]
  87. Tennant, P.F.; Gonsalves, C.; Ling, K.S.; Fitch, M.; Manshardt, R.; Slightom, L.J.; Gonsalves, D. Differential protection against papaya ringspot virus isolates in coat protein gene transgenic papaya and classically cross-protected papaya. Phytopathology 1994, 84, 1359–1366. [Google Scholar] [CrossRef] [Green Version]
  88. Mundembe, R.; Matibiri, A.; Sithole-Niang, I. Transgenic plants expressing the coat protein gene of cowpea aphid-borne mosaic potyvirus predominantly convey the delayed symptom development phenotype. Afr. J. Biotechnol. 2009, 8, 2682–2690. [Google Scholar]
  89. Golemboski, D.B.; Lomonossoff, G.P.; Zaitlin, M. Plants transformed with a tobacco mosaic virus nonstructural gene sequence are resistant to the virus. Proc. Natl. Acad. Sci. USA. 1990, 87, 6311–6315. [Google Scholar] [CrossRef]
  90. Fire, A.; Xu, S.; Montgomery, M.K.; Kostas, S.A.; Driver, S.E.; Mello, C.C. Potent and specific genetic interference by double stranded RNA in Caenorhabditis elegans. Nature 1998, 391, 806–811. [Google Scholar] [CrossRef]
  91. Prins, M.; Laimer, M.; Noris, E.; Schubert, J.; Wassenegger, M.; Tepfer, M. Strategies for antiviral resistance in transgenic plants. Mol. Plant Pathol. 2008, 9, 73–83. [Google Scholar] [CrossRef]
  92. Chen, Z.; Gu, H.; Li, Y.; Su, Y.; Wu, P.; Jiang, Z.; Ming, X.; Tian, J.; Pan, N.; Qu, L.J. Safety assessment for genetically modified sweet pepper and tomato. Toxicology 2003, 188, 297–307. [Google Scholar] [CrossRef]
  93. Tennant, P.; Souza, M.T.; Fitch, M.M.; Manshardt, R.M.; Slightom, J.L. Line 63-1: A new virus-resistant transgenic papaya. Hortscience 2005, 40, 1196–1199. [Google Scholar] [CrossRef] [Green Version]
  94. Wu, Z.; Mo, C.; Zhang, S.; Li, H. Characterization of papaya ringspot virus isolates infecting transgenic papaya ‘Huanong No.1’ in South China. Sci. Rep. 2018, 8, 8206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Brunetti, A.; Tavazza, M.; Noris, E.; Tavazza, R.; Caciagli, P.; Ancora, G.; Crespi, S.; Accotto, G.P. High expression of truncated viral rep protein confers resistance to tomato yellow leaf curl virus in transgenic tomato plants. Mol. Plant Microbe Interact. 1997, 10, 571–579. [Google Scholar] [CrossRef] [Green Version]
  96. Faria, J.C.; Albino, M.M.C.; Dias, B.B.A.; Cançado, L.J.; Cunha, N.B.D.; Silva, L.D.M.; Vianna, G.R.; Aragão, F.J.L. Partial resistance to bean golden mosaic virus in a transgenic common bean (Phaseolus vulgaris L.) line expressing a mutated rep gene. Plant Sci. 2006, 171, 565–571. [Google Scholar] [CrossRef] [Green Version]
  97. Aragao, F.J.L.; Nogueira, E.O.P.L.; Tinoco, M.L.P.; Faria, J.C. Molecular characterization of the first commercial transgenic common bean immune to the bean golden mosaic virus. J. Biotechnol. 2013, 166, 42–50. [Google Scholar] [CrossRef]
  98. Lapidot, M.; Gafny, R.; Ding, B.; Wolf, S.; Lucas, W.J.; Beachy, R.N. A dysfunctional movement protein of tobacco mosaic virus that partially modifies the plasmodesmata and limits virus spread in transgenic plants. Plant J. 1993, 4, 959–970. [Google Scholar] [CrossRef]
  99. Cooper, B.; Lapidot, M.; Heick, J.A.; Dodds, J.A.; Beachy, R.N. A defective movement protein of TMV in transgenic plants confers resistance to multiple viruses whereas the functional analog increases susceptibility. Virology 1995, 206, 307–313. [Google Scholar] [CrossRef]
  100. Carvalho, J.L.; De Oliveira Santos, J.; Conte, C.; Pacheco, S.; Nogueira, E.O.; Souza, T.L.; Faria, J.C.; Aragão, F.J. Comparative analysis of nutritional compositions of transgenic RNAi-mediated virus-resistant bean (event EMB-PV051-1) with its non-transgenic counterpart. Transgenic Res. 2015, 24, 813–819. [Google Scholar] [CrossRef]
  101. Borah, M.; Berbati, M.; Reppa, C.; Holeva, M.; Nath, P.D.; Voloudakis, A. RNA-based vaccination of Bhut Jolokia pepper (Capsicum chinense Jacq.) against cucumber mosaic virus. Virusdisease 2018, 29, 207–211. [Google Scholar] [CrossRef]
  102. Callahan, A.M.; Dardick, C.D.; Scorza, R. Multilocation comparison of fruit composition for ‘HoneySweet’, an RNAi based plum pox virus resistant plum. PLoS ONE 2019, 14, e0213993. [Google Scholar] [CrossRef] [Green Version]
  103. Halterman, D.A.; Kramer, L.C.; Wielgus, S.; Jiang, J. Performance of transgenic potato containing the late blight resistance gene RB. Plant Dis. 2008, 92, 339–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Zaidi, S.S.; Tashkandi, M.; Mansoor, S.; Mahfouz, M.M. Engineering plant immunity: Using CRISPR/Cas9 to generate virus resistance. Front. Plant Sci. 2016, 7, 1673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Khan, M.Z.; Amin, I.; Hameed, A.; Mansoor, S. CRISPR-Cas13a: Prospects for plant virus resistance. Trends Biotechnol. 2018, 36, 1207–1210. [Google Scholar] [CrossRef] [PubMed]
  106. Noureen, A.; Khan, M.Z.; Amin, I.; Zainab, T.; Ahmad, N.; Haider, S.; Mansoor, S. Broad-spectrum resistance against multiple PVY strains by CRSIPR/Cas13 system in Solanum tuberosum crop. GM Crops Food 2022, 13, 97–111. [Google Scholar] [CrossRef] [PubMed]
  107. Stuiver, M.H.; Custers, J.H. Engineering disease resistance in plants. Nature 2001, 411, 865–868. [Google Scholar] [CrossRef] [Green Version]
  108. Wally, O.; Punja, Z.K. Genetic engineering for increasing fungal and bacterial disease resistance in crop plants. GM Crops 2010, 1, 199–206. [Google Scholar] [CrossRef]
  109. Hwang, B.H.; Bae, H.; Lim, H.S.; Kim, K.B.; Kim, S.J.; Im, M.H.; Park, B.S.; Kim, J. Overexpression of polygalacturonase-inhibiting protein 2 (PGIP2) of Chinese cabbage (Brassica rapa ssp. pekinensis) increased resistance to the bacterial pathogen Pectobacterium carotovorum ssp. carotovorum. Plant Cell Tissue Organ Cult. 2010, 103, 293–305. [Google Scholar] [CrossRef]
  110. Wang, Z.; Fang, H.; Chen, Y.; Chen, K.; Li, G.; Gu, S.; Tan, X. Overexpression of BnWRKY33 in oilseed rape enhances resistance to Sclerotinia sclerotiorum. Mol. Plant Pathol. 2014, 15, 677–689. [Google Scholar] [CrossRef]
  111. Zhang, Y.; Huai, D.; Yang, Q.; Cheng, Y.; Ma, M.; Kliebenstein, D.J.; Zhou, Y. Overexpression of three glucosinolate iosynthesis genes in Brassica napus identifies enhanced resistance to Sclerotinia sclerotiorum and Botrytis cinerea. PLoS ONE 2015, 10, e0140491. [Google Scholar]
  112. Fang, Y.; Tyler, B.M. Efficient disruption and replacement of an effector gene in the oomycete Phytophthora sojae using CRISPR/Cas9. Mol. Plant Pathol. 2016, 17, 127–139. [Google Scholar] [CrossRef]
  113. Ali, S.; Mir, Z.A.; Tyagi, A.; Mehari, H.; Meena, R.P.; Bhat, A.; Yadav, P.; Papalou, P.; Rawat, S.; Grover, A. Overexpression of NPR1 in Brassica juncea confers road spectrum resistance to fungal pathogens. Front. Plant Sci. 2017, 8, 1693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Cui, J.; Jiang, N.; Zhou, X.; Hou, X.; Yang, G.; Meng, J.; Luan, Y. Tomato MYB49 enhances resistance to Phytophthora infestans and tolerance to water deficit and salt stress. Planta 2018, 248, 1487–1503. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, T.; Chen, T.; Kan, J.; Yao, Y.; Guo, D.; Yang, Y.; Ling, X.; Wang, J.; Zhang, B. The GhMYB36 transcription factor confers resistance to biotic and abiotic stress by enhancing PR1 gene expression in plants. Plant Biotechnol. J. 2022, 20, 722–735. [Google Scholar] [CrossRef] [PubMed]
  116. Anand, A.; Zhou, T.; Trick, H.N.; Gill, B.S.; Bockus, W.W.; Muthukrishnan, S. Greenhouse and field testing of transgenic wheat plants stably expressing genes for thaumatin-like protein, chitinase and glucanase against Fusarium graminearum. J. Exp. Bot. 2003, 54, 1101–1111. [Google Scholar] [CrossRef] [Green Version]
  117. Yang, S.; Gao, M.; Xu, C.; Gao, J.; Deshpande, S.; Lin, S.; Roe, B.A.; Zhu, H. Alfalfa benefits from Medicago truncatula: The RCT1 gene from M. truncatula confers broad spectrum resistance to anthracnose in alfalfa. Proc. Natl. Acad. Sci. USA 2008, 105, 12164–12169. [Google Scholar] [CrossRef] [Green Version]
  118. Alexander, D.; Goodman, R.M.; Gut-Rella, M.; Glascock, C.; Weymann, K.; Friedrich, L.; Maddox, D.; Ahl-Goy, P.; Luntz, T.; Ward, E.; et al. Increased tolerance to two oomycete pathogens in transgenic tobacco expressing pathogenesis-related protein 1a. Proc. Natl. Acad. Sci. USA 1993, 90, 7327–7331. [Google Scholar] [CrossRef] [Green Version]
  119. Doares, S.H.; Narvaez-Vazquez, J.; Conconi, A.; Ryan, C.A. Salicylic acid inhibits synthesis of proteinase inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 1995, 108, 1741–1746. [Google Scholar] [CrossRef]
  120. 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]
  121. Song, Y.; Liu, L.; Wang, Y.; Valkenburg, D.J.; Zhang, X.; Zhu, L.; Thomma, B.P.H.J. Transfer of tomato immune receptor Ve1 confers Ave1-dependent Verticillium resistance in tobacco and cotton. Plant Biotechnol. J. 2018, 16, 638–648. [Google Scholar] [CrossRef] [Green Version]
  122. Du, J.; Verzaux, E.; Chaparro-Garcia, A.; Bijsterbosch, G.; Keizer, L.C.P.; Zhou, J.; Liebrand, T.W.H.; Xie, C.; Govers, F.; Robatzek, S.; et al. Elicitin recognition confers enhanced resistance to Phytophthora infestans in potato. Nat. Plants 2015, 1, 15034. [Google Scholar] [CrossRef]
  123. Foster, S.J.; Park, T.H.; Pel, M.; Brigneti, G.; Sliwka, J.; Jagger, L.; van der Vossen, E.; Jones, J.D. Rpi-vnt1.1, a Tm-2(2) homolog from Solanum venturii, confers resistance to potato late blight. Mol. Plant Microbe Interact. 2009, 22, 589–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Luo, M.L.; Xie, S.; Chakraborty, A.; Wang, O.; Matny, M.; Jugovich, J.A.; Kolmer, T.; Richardson, D.; Bhatt, M.; Hoque, M.; et al. A five-transgene cassette confers broad-spectrum resistance to a fungal rust pathogen in wheat. Nat. Biotechnol. 2021, 39, 561–566. [Google Scholar] [CrossRef] [PubMed]
  125. Mourgues, F.; Brisset, M.N.; Chevreau, E. Strategies to improve plant resistance to bacterial diseases through genetic engineering. Trends Biotechnol. 1998, 16, 203–210. [Google Scholar] [CrossRef]
  126. Zhou, Y.L.; Xu, J.L.; Zhou, S.C.; Yu, J.; Xie, X.W.; Xu, M.R.; Sun, Y.; Zhu, L.H.; Fu, B.Y.; Gao, Y.M.; et al. Pyramiding Xa23 and Rxo1 for resistance to two bacterial diseases into an elite indica rice variety using molecular approaches. Mol. Breed. 2009, 23, 279–287. [Google Scholar] [CrossRef]
  127. Jaynes, J.M.; Nagpala, P.; Destéfano-Beltrán, L.; Huang, J.H.; Kim, J.H.; Denny, T.; Cetiner, S. Expression of a cecropin B lytic peptide analog in transgenic tobacco confers enhanced resistance to bacterial wilt caused by Pseudomonas solanacearum. Plant Sci. 1993, 89, 43–53. [Google Scholar] [CrossRef]
  128. Huang, Y.; Nordeen, R.O.M.D.; Di, M.; Owens, L.D.; Mcbeath, J.H. Expression of an engineered cecropin gene cassette in transgenic tobacco plants confers disease resistance to Pseudomonas syringae pv. tabaci. Phytopathology 1997, 87, 494–499. [Google Scholar] [CrossRef] [Green Version]
  129. Mitra, A.; Zhang, Z. Expression of a human lactoferrin cDNA in tobacco cells produces antibacterial protein(s). Plant Physiol. 1994, 106, 977–981. [Google Scholar] [CrossRef] [Green Version]
  130. Trudel, J.; Potvin, C.; Asselin, A. Secreted hen lysozyme in transgenic tobacco: Recovery of bound enzyme and in vitro growth inhibition of plant pathogens. Plant Sci. 1995, 106, 55–62. [Google Scholar] [CrossRef]
  131. Nakajima, H.; Muranaka, T.; Ishige, F.; Akutsu, K.; Oeda, K. Fungal and bacterial disease resistance in transgenic plants expressing human lysozyme. Plant Cell Rep. 1997, 16, 674–679. [Google Scholar] [CrossRef]
  132. Zhang, Z.; Coyne, D.P.; Vidaver, A.K.; Mitra, A. Expression of human lactoferrin cDNA confers resistance to Ralstonia solanacearum in transgenic tobacco plants. Phytopathology 1998, 88, 730–734. [Google Scholar] [CrossRef] [Green Version]
  133. Rivero, M.; Furman, N.; Mencacci, N.; Picca, P.; Toum, L.; Lentz, E.; Bravo-Almonacid, F.; Mentaberry, A. Stacking of antimicrobial genes in potato transgenic plants confers increased resistance to bacterial and fungal pathogens. J. Biotechnol. 2012, 157, 334–343. [Google Scholar] [CrossRef] [PubMed]
  134. Lakshman, D.K.; Natarajan, S.; Mandal, S.; Mitra, A. Lactoferrin-derived resistance against plant pathogens in transgenic plants. J. Agric. Food Chem. 2013, 61, 11730–11735. [Google Scholar] [CrossRef] [PubMed]
  135. Maleck, K.; Levine, A.; Eulgem, T.; Morgan, A.; Schmid, J.; Lawton, K.A.; Dangl, J.L.; Dietrich, R.A. The transcriptome of Arabidopsis thaliana during systemic acquired resistance. Nat. Genet. 2000, 26, 403–410. [Google Scholar] [CrossRef] [PubMed]
  136. Zipfel, C.; Kunze, G.; Chinchilla, D.; Caniard, A.; Jones, J.D.G.; Boller, T.; Felix, G. Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts agrobacterium-mediated transformation. Cell 2006, 125, 749–760. [Google Scholar] [CrossRef] [PubMed]
  137. Lacombe, S.; Rougon-Cardoso, A.; Sherwood, E.; Peeters, N.; Dahlbeck, D.; van Esse, H.P.; Smoker, M.; Rallapalli, G.; Thomma, B.P.H.J.; Staskawicz, B.; et al. Interfamily transfer of a plant pattern-recognition receptor confers broad-spectrum bacterial resistance. Nat. Biotechnol. 2010, 28, 365–369. [Google Scholar] [CrossRef] [PubMed]
  138. Schwessinger, B.; Bahar, O.; Thomas, N.; Holton, N.; Nekrasov, V.; Ruan, D.; Canlas, P.E.; Daudi, A.; Petzold, C.J.; 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]
  139. Schoonbeek, H.; Wang, H.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]
  140. 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: EFR confers bacterial disease resistance in monocot rice. J. Integr. Plant. Biol. 2015, 57, 641–652. [Google Scholar] [CrossRef]
  141. Boschi, F.; Schvartzman, C.; Murchio, S.; Ferreira, V.; Siri, M.I.; Galván, G.A.; Smoker, M.; Stransfeld, L.; Zipfel, C.; Vilaró, F.L.; et al. Enhanced bacterial wilt resistance in potato through expression of Arabidopsis EFR and introgression of quantitative resistance from Solanum commersonii. Front. Plant Sci. 2017, 8, 1642. [Google Scholar] [CrossRef]
  142. Pfeilmeier, S.; George, J.; Morel, A.; Roy, S.; Smoker, M.; Stransfeld, L.; Downie, J.A.; Peeters, N.; Malone, J.G.; Zipfel, C. Expression of the Arabidopsis thaliana immune receptor EFR in Medicago truncatula reduces infection by a root pathogenic bacterium, but not nitrogen-fixing rhizobial symbiosis. Plant Biotechnol. J. 2019, 17, 569–579. [Google Scholar] [CrossRef] [Green Version]
  143. Lilley, C.J.; Devlin, F.; Urwin, P.E.; Atkinson, H.J. Parasitic nematodes, proteinases and transgenic plants. Parasitol. Today 1999, 15, 414–417. [Google Scholar] [CrossRef]
  144. Bakhetia, M.; Charlton, W.L.; Urwin, P.E.; McPherson, M.J.; Atkinson, H.J. RNA interference and plant parasitic nematodes. Trends Plant Sci. 2005, 10, 362–367. [Google Scholar] [CrossRef] [PubMed]
  145. Rosso, M.N.; Dubrana, M.P.; Cimbolini, N.; Jaubert, S.; Abad, P. Application of RNA interference to root-knot nematode genes encoding esophageal gland proteins. Mol. Plant Microbe Interact. 2005, 18, 615–620. [Google Scholar] [CrossRef] [Green Version]
  146. Gheysen, G.; Kyndt, T.; Haegeman, A.; Joseph, S.; Remy, S.; Swennen, R. RNAi for research and applications in plant-nematode interactions. In Vitro Cell. Dev. Biol. Anim. 2010, 46, S28–S29. [Google Scholar]
  147. Bakhetia, M.; Charlton, W.; Atkinson, H.J.; McPherson, M.J. RNA interference of dual oxidase in the plant nematode Meloidogyne incognita. Mol. Plant Microb. Interact. 2005, 18, 1099–1106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Yadav, B.C.; Veluthambi, K.; Subramaniam, K. Host generated double stranded RNA induces RNAi in plant parasitic nematodes and protects the host from infection. Mol. Biochem. Parasitol. 2006, 148, 219–222. [Google Scholar] [CrossRef]
  149. Shingles, J.; Lilley, C.J.; Atkinson, H.J.; Urwin, P.E. Meloidogyne incognita: Molecular and biochemical characterisation of a cathepsin L cysteine proteinase and the effect on parasitism following RNAi. Exp. Parasitol. 2007, 115, 114–120. [Google Scholar] [CrossRef]
  150. Charlton, W.L.; Harel, H.Y.M.; Bakhetia, M.; Hibbard, J.K.; Atkinson, H.J.; Mcpherson, M.J. Additive effects of plant expressed double-stranded RNAs on root-knot nematode development. Int. J. Parasitol. 2010, 40, 855–864. [Google Scholar] [CrossRef]
  151. Ibrahim, H.M.; Alkharouf, N.W.; Meyer, S.L.; Aly, M.A.; Gamal, E.A.K. Post-transcriptional gene silencing of root-knot nematode in transformed soybean roots. Exp. Parasitol. 2011, 127, 90–99. [Google Scholar] [CrossRef]
  152. Niu, J.H.; Jian, H.; Xu, J.; Chen, C.; Guo, Q. RNAi silencing of the Meloidogyne incognita Rpn7 gene reduces nematode parasitic success. Euro. J. Plant Pathol. 2012, 134, 131–144. [Google Scholar] [CrossRef]
  153. Papolu, P.K.; Gantasala, N.P.; Kamaraju, D.; Banakar, P.; Sreevathsa, R.; Rao, U. Utility of host delivered RNAi of two FMRF amide like peptides, flp-14 and flp-18, for the management of root-knot nematode, Meloidogyne incognita. PLoS ONE 2013, 8, e80603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Antonino de Souza, J.D., Jr.; Ramos Coelho, R.; Tristan Lourenço, I.; da Rocha Fragoso, R.; Barbosa Viana, A.A.; Lima Pepino de Macedo, L.; Mattar da Silva, M.C.; Gomes Carneiro, R.M.; Engler, G.; de Almeida-Engler, J.; et al. Knocking- down Meloidogyne incognita proteases by plant-delivered dsRNA has negative pleiotropic effect on nematode vigor. PLoS ONE 2013, 8, e85364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Dutta, T.K.; Papolu, P.K.; Banakar, P.; Choudhary, D.; Sirohi, A.; Rao, U. Tomato transgenic plants expressing hairpin construct of a nematode protease gene conferred enhanced resistance to root-knot nematodes. Front. Microbiol. 2015, 6, 260. [Google Scholar] [CrossRef] [PubMed]
  156. Dutta, T.K.; Banakar, P.; Rao, U. The status of RNAi-based transgenic research in plant nematology. Front. Microbiol. 2015, 5, 760. [Google Scholar] [CrossRef]
  157. Walawage, S.L.; Britton, M.T.; Leslie, C.A.; Uratsu, S.L.; Li, Y.; Dandekar, M. Stacking resistance to crown gall and nematodes in walnut rootstocks. BMC Genomics 2013, 14, 668. [Google Scholar] [CrossRef] [Green Version]
  158. Dinh, P.T.Y.; Zhang, L.; Brown, C.R.; Elling, A.A. Plant-mediated RNA interference of effector gene Mc16D10L confers resistance against Meloidogyne chitwoodi in diverse genetic backgrounds of potato and reduces pathogenicity of nematode offspring. Nematology 2014, 6, 669–682. [Google Scholar] [CrossRef]
  159. Klink, V.P.; Kim, K.H.; Martins, V.; Macdonald, M.H.; Beard, H.S.; Alkharouf, N.W.; Lee, S.K.; Park, S.C.; Matthews, B.F. A correlation between host-mediated expression of parasite genes as tandem inverted repeats and abrogation of development of female Heterodera glycines cyst formation during infection of Glycine max. Planta 2009, 230, 53–71. [Google Scholar] [CrossRef]
  160. Li, J.; Todd, T.C.; Oakley, T.R.; Lee, J.L.; Trick, H.N. Host-derived suppression of nematode reproductive and fitness genes decreases fecundity of Heterodera glycines Ichinohe. Planta 2010, 232, 775–785. [Google Scholar] [CrossRef]
  161. Danchin, E.G.J.; Arguel, M.J.; Campan-Fournier, A.; Perfus-Barbeoch, L.; Magliano, M.; Rosso, M.N.; Da Rocha, M.; Da Silva, C.; Nottet, N.; Labadie, K.; et al. Identification of novel target genes for safer and more specific control of root-knot nematodes from a pan-genome mining. PLoS Pathogen. 2013, 9, e1003745. [Google Scholar] [CrossRef] [Green Version]
  162. Yu, H.; Yang, Q.; Fu, F.; Li, W. Three strategies of transgenic manipulation for crop improvement. Front. Plant Sci. 2022, 13, 948518. [Google Scholar] [CrossRef]
  163. Liu, D. Design of gene constructs for transgenic maize. Meth. Mol. Biol. 2009, 526, 3–20. [Google Scholar]
  164. Parvathy, S.T.; Udayasuriyan, V.; Bhadana, V. Codon usage bias. Mol. Biol. Rep. 2022, 49, 539–565. [Google Scholar] [CrossRef] [PubMed]
  165. Adamczyk, J.J.; Meredith, W.R. Genetic basis for variability of Cry1Ac expression among commercial transgenic Bacillus thuringiensis (Bt) cotton cultivars in the United States. J. Cotton Sci. 2004, 8, 17–23. [Google Scholar]
  166. Poongothai, S.; Iiavarasan, R.; Karrunakaran, C.M. Cry1Ac levels and biochemical variations in Bt cotton as influenced by tissue maturity and senescence. J. Plant Breed. Crop Sci. 2010, 2, 96–103. [Google Scholar]
  167. Chen, Y.; Li, Y.; Zhou, M.; Cai, Z.; Tambel, L.I.M.; Zhang, X.; Chen, Y.; Chen, D. Nitrogen deficit decreases seed Cry1Ac endotoxin expression in Bt transgenic cotton. Plant Physiol. Biochem. 2019, 141, 114–121. [Google Scholar] [CrossRef] [PubMed]
  168. Adamczyk, J.J.; Sumerford, D.V. Potential factors impacting season-long expression of Cry1Ac in 13 commercial varieties of Bollgard cotton. J. Insect Sci. 2001, 1, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Adamczyk, J.J.; Hardee, D.D.; Adams, L.C.; Sumerford, D.V. Correlating differences in larval survival and development of bollworms (Lepidoptera: Noctuidae) and fall armyworms (Lepidoptera: Noctuidae) to differential expression of Cry1A(c) δ–endotoxin in various plant parts among commercial cultivars of transgenic Bacillus thuringiensis cotton. J. Econ. Entomol. 2001, 94, 284–290. [Google Scholar] [PubMed]
  170. Adamczyk, J.J.; Perera, O.; Meredith, W.R. Production of mRNA from the Cry1Ac transgene differs among Bollgard lines which correlates to the level of subsequent protein. Transgenic Res. 2009, 18, 143–149. [Google Scholar] [CrossRef] [PubMed]
  171. Zhang, L.; Shen, W.; Fang, Z.; Liu, B. Cry1ab/c in different stages of growth in transgenic rice Bt-shanyou63. Front. Biosci. 2016, 21, 447–454. [Google Scholar]
  172. Sanahuja, G.; Banakar, R.; Twyman, R.M.; Capell, T.; Christou, P. Bacillus thuringiensis: A century of research, development and commercial applications. Plant Biotechnol. J. 2011, 9, 283–300. [Google Scholar] [CrossRef] [Green Version]
  173. Palma, L.; Muñoz, D.; Berry, C.; Murillo, J.; Caballero, P. Bacillus thuringiensis toxins: An overview of their biocidal activity. Toxins 2014, 6, 3296–3325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Melo, A.L.; Soccol, V.T.; Soccol, C.R. Bacillus thuringiensis: Mechanism of action, resistance, and new applications: A review. Crit. Rev. Biotechnol. 2016, 36, 317–326. [Google Scholar] [CrossRef] [PubMed]
  175. de Wit, P.J.G.M. Molecular characterization of gene-for-gene systems in plant-fungus interactions and the application of avirulence genes in control of plant pathogens. Annu. Rev. Phytopathol. 1992, 30, 391–418. [Google Scholar] [CrossRef]
  176. Kamoun, S.; van West, P.; Vleeshouwers, V.G.A.A.; de Groot, K.E.; Govers, F. Resistance of Nicotiana benthamiana to Phytophthora infestans is mediated by the recognition of the elicitor protein INF1. Plant Cell 1998, 10, 1413–1425. [Google Scholar] [CrossRef] [PubMed]
  177. Keller, H.; Pamboukdjian, N.; Ponchet, M.; Poupet, A.; Delon, R.; Verrier, J.L.; Roby, D.; Ricci, P. Pathogen induced elicitin production in transgenic tobacco generates a hypersensitive response and nonspecific disease resistance. Plant Cell 1999, 11, 223–235. [Google Scholar] [CrossRef] [Green Version]
  178. Heath, M.C. Nonhost resistance and non-specific plant defenses. Curr. Opin. Plant Biol. 2000, 3, 315–319. [Google Scholar] [CrossRef]
  179. Stalker, D.M.; Mcbride, K.E.; Malyj, L.D. Herbicide resistance in transgenic plants expressing a bacterial detoxification gene. Science 1988, 242, 419–423. [Google Scholar] [CrossRef]
  180. Matten, S.R.; Reynolds, A.H. Current resistance management requirements for Bt cotton in the United States. J. New Seeds 2003, 5, 137–178. [Google Scholar] [CrossRef]
  181. Fraser, M.J., Jr. Insect transgenesis: Current applications and future prospects. Annu. Rev. Entomol. 2012, 57, 267–289. [Google Scholar] [CrossRef] [Green Version]
  182. Davis, M.J.; Ying, Z. Development of papaya breeding lines with transgenic resistance to papaya ringspot virus. Plant Dis. 2004, 88, 352–358. [Google Scholar] [CrossRef] [Green Version]
  183. Ilardi, V.; Nicola-Negri, E.D. Genetically engineered resistance to plum pox virus infection in herbaceous and stone fruit hosts. GM Crop 2011, 2, 24–33. [Google Scholar] [CrossRef]
  184. Orbegozo, J.; Solorzano, D.; Cuellar, W.J.; Bartolini, I.; Roman, M.L.; Ghislain, M.; Kreuze, J. Marker-free PLRV resistant potato mediated by Cre-loxP excision and RNAi. Transgenic Res. 2016, 25, 813–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Chiozza, M.V.; Burachik, M.; Miranda, P.V. Compositional analysis of soybean event IND–ØØ41Ø–5. GM Crops Food 2020, 11, 154–163. [Google Scholar] [CrossRef]
  186. Demaneche, S.; Sanguin, H.; Pote, J.; Navarro, E.; Bernillon, D.; Mavingui, P.; Wildi, W.; Vogel, T.M.; Simonet, P. Antibiotic-resistant soil bacteria in transgenic plant fields. Proc. Natl. Acad. Sci. USA 2008, 105, 3957–3962. [Google Scholar] [CrossRef] [PubMed]
  187. Song, J.; Bradeen, J.M.; Naess, S.K.; Raasch, J.A.; Wielgus, S.M.; Haberlach, G.T.; Liu, J.; Kuang, H.; Austin-Phillips, S.; Buell, C.R.; et al. Gene RB cloned from Solanum bulbocastanum confers broad spectrum resistance to potato late blight. Proc. Natl. Acad. Sci. USA. 2003, 100, 9128–9133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Yin, Z.; Hennig, J.; Szwacka, M.; Malepszy, S. Tobacco PR-2d promoter is induced in transgenic cucumber in response to biotic and abiotic stimuli. J. Plant Physiol. 2004, 161, 621–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Kim, Y.S.; Lim, S.; Yoda, H.; Choi, Y.E.; Sano, H. Simultaneous activation of salicylate production and fungal resistance in transgenic Chrysanthemum producing caffeine. Plant Signal Behav. 2011, 6, 409–412. [Google Scholar] [CrossRef] [Green Version]
  190. Jisha, V.; Dampanaboina, L.; Vadassery, J.; Mithöfer, A.; Kappara, S.; Ramanan, R. Overexpression of an AP2/ERF type transcription factor OsEREBP1 confers biotic and abiotic stress tolerance in rice. PLoS ONE 2015, 10, e0127831. [Google Scholar] [CrossRef]
  191. Coppola, M.; Corrado, G.; Coppola, V.; Cascone, P.; Martinelli, R.; Digilio, M.C.; Pennacchio, F.; Rao, R. Prosystemin overexpression in tomato enhances resistance to different biotic stresses by activating genes of multiple signaling pathways. Plant Mol. Biol. Rep. 2015, 33, 1270–1285. [Google Scholar] [CrossRef] [Green Version]
  192. Liu, Q.; Liu, Y.; Tang, Y.; Chen, J.; Ding, W. Overexpression of NtWRKY50 increases resistance to Ralstonia solanacearum and alters salicylic acid and jasmonic acid production in tobacco. Front. Plant Sci. 2017, 8, 1710. [Google Scholar] [CrossRef] [Green Version]
  193. Klee, H.J. Ripening physiology of fruit from transgenic tomato (Lycopersicon esculentum) plants with reduced ethylene synthesis. Plant Physiol. 1993, 102, 911–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Taning, C.N.T.; Arpaia, S.; Christiaens, O.; Dietz-Pfeilstetter, A.; Jones, H.; Mezzetti, B.; Sabbadini, S.; Sorteberg, H.G.; Sweet, J.; Ventura, V.; et al. RNA-based biocontrol compounds: Current status and perspectives to reach the market. Pest Manag. Sci. 2020, 76, 841–845. [Google Scholar] [CrossRef] [PubMed]
  195. Giudice, G.; Moffa, L.; Varotto, S.; Cardone, M.F.; Bergamini, C.; de Lorenzis, G.; Velasco, R.; Nerva, L.; Chitarra, W. Novel and emerging biotechnological crop protection approaches. Plant Biotechnol. J. 2021, 19, 1495–1510. [Google Scholar] [CrossRef] [PubMed]
  196. Rajam, M.V. RNA silencing technology: A boon for crop improvement. J. Biosci. 2020, 45, 118. [Google Scholar] [CrossRef] [PubMed]
  197. Mitter, N.; Worrall, E.A.; Robinson, K.E.; Li, P.; Jain, R.G.; Taochy, C.; Fletcher, S.J.; Carroll, B.J.; Lu, G.Q.; Xu, Z.P. Clay nanosheets for topical delivery of RNAi for sustained protection against plant viruses. Nat. Plants 2017, 3, 16207. [Google Scholar] [CrossRef]
  198. Senthil-Kumar, M.; Mysore, K.S. Caveat of RNAi in plants: The off-target effect. Methods Mol. Biol. 2011, 744, 13–25. [Google Scholar]
  199. Neumeier, J.; Meister, G. siRNA Specificity: RNAi mechanisms and strategies to reduce off-target effects. Front. Plant Sci. 2021, 11, 526455. [Google Scholar] [CrossRef]
  200. Chen, J.; Peng, Y.; Zhang, H.; Wang, K.; Zhao, C.; Zhu, G.; Reddy Palli, S.; Han, Z. Off-target effects of RNAi correlate with the mismatch rate between dsRNA and non-target mRNA. RNA Biol. 2021, 18, 1747–1759. [Google Scholar] [CrossRef]
  201. Pandey, P.; Mysore, K.S.; Senthil-Kumar, M. Recent advances in plant gene silencing methods. Methods Mol. Biol. 2022, 2408, 1–22. [Google Scholar]
  202. Proctor, R.H.; Hohn, T.M.; McCormick, S.P. Reduced virulence of Gibberella zeae caused by disruption of a trichothecene toxin biosynthetic gene. Mol. Plant-Microbe Interact. 1995, 8, 593–601. [Google Scholar] [CrossRef] [Green Version]
  203. Daub, M.E.; Ehrenshaft, M. The photoactivated Cercospora toxin cercosporin: Contributions to plant disease and fundamental biology. Annu. Rev. Phytopathol. 2000, 38, 461–490. [Google Scholar] [CrossRef] [Green Version]
  204. Cessna, S.G.; Sear, V.E.; Dickman, M.B.; Low, P.S. Oxalic acid, a pathogenicity factor for Sclerotinia sclerotiorum, suppresses the oxidative burst of the host plant. Plant Cell 2000, 12, 2191–2200. [Google Scholar] [CrossRef] [Green Version]
  205. Gaj, T.; Gersbach, C.A.; Barbas, C.F., III. ZFN, TALEN, and CRISPR/Cas–based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [Green Version]
  206. Leibowitz, M.L.; Papathanasiou, S.; Doerfler, P.A.; Blaine, L.J.; Sun, L.; Yao, Y.; Zhang, C.Z.; Weiss, M.J.; Pellman, D. Chromothripsis as an on-target consequence of CRISPR-Cas9 genome editing. Nat. Genet. 2021, 53, 889–905. [Google Scholar] [CrossRef] [PubMed]
  207. Molla, K.A.; Sretenovic, S.; Bansal, K.C.; Qi, Y. Precise plant genome editing using base editors and prime editors. Nat. Plants 2021, 7, 1166–1187. [Google Scholar] [CrossRef] [PubMed]
  208. Nerkar, G.; Devarumath, S.; Purankar, M.; Kumar, A.; Valarmathi, R.; Devarumath, R.; Appunu, C. Advances in crop breeding through precision genome editing. Front. Genet. 2022, 13, 880195. [Google Scholar] [CrossRef] [PubMed]
  209. Chilcoat, D.; Liu, Z.B.; Sander, J. Use of CRISPR/Cas9 for crop improvement in maize and soybean. Prog. Mol. Biol. Transl. Sci. 2017, 149, 27–46. [Google Scholar]
  210. Sánchez-León, S.; Gil-Humanes, J.; Ozuna, C.V.; Giménez, M.J.; Sousa, C.; Voytas, D.F.; Barro, F. Low-gluten, nontransgenic wheat engineered with CRISPR/Cas9. Plant Biotechnol. J. 2018, 16, 902–910. [Google Scholar] [CrossRef] [PubMed]
  211. Nekrasov, V.; Staskawicz, B.; Weigel, D.; Jones, D.; Kamoun, S. Targeted mutagenesis in the model plant Nicotiana enthamiana using Cas9 RNA–guided endonuclease. Nat. Biotechnol. 2013, 31, 691–693. [Google Scholar] [CrossRef]
  212. Shan, Q.; Wang, Y.; Li, J.; Zhang, Y.; Chen, K.; Liang, Z.; Zhang, K.; Liu, J.; Xi, J.J.; Qiu, J.L.; et al. Targeted genome modification of crop plants using a CRISPR/Cas system. Nat. Biotechnol. 2013, 31, 686–688. [Google Scholar] [CrossRef]
  213. van der Oost, J.; Westra, E.R.; Jackson, R.N.; Wiedenheft, B. Unravelling the structural and mechanistic basis of CRISPR–Cas systems. Nat Rev Microbiol. 2014, 12, 479–492. [Google Scholar] [CrossRef] [PubMed]
  214. Lawrenson, T.; Shorinola, O.; Stacey, N.; Li, C.; Østergaard, L.; Patron, N.; Uauy, C.; Harwood, W. Induction of targeted, heritable mutations in barley and Brassica oleracea using RNA–guided Cas9 nuclease. Genome Biol. 2015, 16, 258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Malnoy, M.; Viola, R.; Jung, M.H.; Koo, O.J.; Kim, S.; Kim, J.S.; Velasco, R.; Nagamangala Kanchiswamy, C.D. DNA-free genetically edited grapevine and apple protoplast using CRISPR/Cas9 ribonucleoproteins. Front. Plant Sci. 2016, 7, 1904. [Google Scholar] [CrossRef] [PubMed]
  216. Hu, X.; Wang, C.; Fu, Y.; Liu, Q.; Jiao, X.; Wang, K. Expanding the range of CRISPR/Cas9 genome editing in rice. Mol. Plant 2016, 9, 943–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Naeem, M.; Majeed, S.; Hoque, M.Z.; Ahmad, I.J.C. Latest developed strategies to minimize the off-target effects in CRISPR-Cas mediated genome editing. Cells 2020, 9, 1608. [Google Scholar] [CrossRef] [PubMed]
  218. Honée, G. Engineered resistance against fungal plant pathogens. Eur. J. Plant Pathol. 1999, 105, 319–326. [Google Scholar] [CrossRef]
  219. Zhang, H.; Zhang, J.; Wei, P.; Zhang, B.; Gou, F.; Feng, Z.; Mao, Y.; Yang, L.; Zhang, H.; Xu, N.; et al. The CRISPR/Cas9 system produces specific and homozygous targeted gene editing in rice in one generation. Plant Biotechnol. J. 2014, 12, 797–807. [Google Scholar] [CrossRef] [PubMed]
  220. 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]
  221. Zhu, J.; Song, N.; Sun, S.; Yang, W.; Zhao, H.; Song, W.; Lai, J. Efficiency and inheritance of targeted mutagenesis in maize using CRISPR-Cas9. J. Genet. Genom. 2016, 43, 25–36. [Google Scholar] [CrossRef]
  222. Waltz, E. Gene-edited CRISPR mushroom escapes US regulation. Nature 2016, 532, 293. [Google Scholar] [CrossRef] [Green Version]
  223. 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] [PubMed] [Green Version]
  224. Waltz, E. CRISPR-edited crops free to enter market, skip regulation. Nat. Biotechnol. 2016, 34, 582. [Google Scholar] [CrossRef] [PubMed]
  225. Shimatani, Z.; Kashojiya, S.; Takayama, M.; Terada, R.; Arazoe, T.; Ishii, H.; Teramura, H.; Yamamoto, T.; Komatsu, H.; Miura, K.; et al. Targeted base editing in rice and tomato using a CRISPR/Cas9 cytidine deaminase fusion. Nat. Biotechnol. 2017, 35, 441–443. [Google Scholar] [CrossRef] [PubMed]
  226. Braatz, J.; Harloff, H.J.; Mascher, M.; Stein, N.; Himmelbach, A.; Jung, C. CRISPR-Cas9 targeted mutagenesis leads to simultaneous modification of different homoeologous gene copies in polyploid oilseed rape (Brassica napus). Plant Physiol. 2017, 174, 935–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Okuzaki, A.; Ogawa, T.; Koizuka, C.; Kaneko, K.; Inaba, M.; Imamura, J.; Koizuka, N. CRISPR/Cas9-mediated genome editing of the fatty acid desaturase 2 gene in Brassica napus. Plant Physiol. Biochem. 2018, 131, 63–69. [Google Scholar] [CrossRef]
  228. Langner, T.; Kamoun, S.; Belhaj, K. CRISPR crops: Plant genome editing toward disease resistance. Ann. Rev. Phytopathol. 2018, 56, 479–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Zhang, Z.; Ge, X.; Luo, X.; Wang, P.; Fan, Q.; Hu, G.; Xiao, J.; Li, F.; Wu, J. Simultaneous editing of two copies of Gh14–3–3d confers enhanced transgene-clean plant defense against Verticillium dahliae in allotetraploid upland cotton. Front. Plant Sci. 2018, 9, 842. [Google Scholar] [CrossRef]
  230. 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] [CrossRef]
  231. Wang, Y.; Wang, J.; Guo, S.; Tian, S.; Zhang, J.; Ren, Y.; Li, M.; Gong, G.; Zhang, H.; Xu, Y. CRISPR/Cas9-mediated mutagenesis of ClBG1 decreased seed size and promoted seed germination in watermelon. Hortic. Res. 2021, 8, 1–12. [Google Scholar] [CrossRef]
  232. Hinge, V.R.; Chavhan, R.L.; Kale, S.P.; Suprasanna, P.; Kadam, U.S. Engineering resistance against viruses in field crops using CRISPR-Cas9. Curr. Genom. 2021, 22, 214–231. [Google Scholar] [CrossRef]
  233. Usman, B.; Nawaz, G.; Zhao, N.; Liao, S.; Qin, B.; Liu, F.; Liu, Y.; Li, R. Programmed editing of rice (Oryza sativa L.) OsSPL16 gene using CRISPR/Cas9 improves grain yield by modulating the expression of pyruvate enzymes and cell cycle proteins. Int. J. Mol. Sci. 2021, 22, 249. [Google Scholar] [CrossRef] [PubMed]
  234. Zhou, X.; Liao, H.; Chern, M.; Yin, J.; Chen, Y.; Wang, J.; Zhu, X.; Chen, Z.; Yuan, C.; Zhao, W.; et al. Loss of function of a rice TPR-domain RNA-binding protein confers broad-spectrum disease resistance. Proc. Natl. Acad. Sci. USA 2018, 115, 3174–3179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Oliva, R.; Ji, C.; Atienza-Grande, G.; Huguet-Tapia, J.C.; Perez-Quintero, A.; Li, T.; Eom, J.S.; Li, C.; Nguyen, H.; Liu, B.; et al. Broad-spectrum resistance to bacterial blight in rice using genome editing. Nat. Biotechnol. 2019, 37, 1344–1350. [Google Scholar] [CrossRef] [Green Version]
  236. 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] [PubMed] [Green Version]
  237. Gomez, M.A.; Lin, Z.D.; Moll, T.; Chauhan, R.D.; Hayden, L.; Renninger, K.; Beyene, G.; Taylor, N.J.; Carrington, J.C.; Staskawicz, B.J.; et al. Simultaneous CRISPR/Cas9-mediated editing of cassava eIF4E isoforms nCBP-1 and nCBP-2 reduces cassava brown streak disease symptom severity and incidence. Plant Biotechnol. J. 2019, 17, 421–434. [Google Scholar] [CrossRef] [Green Version]
  238. Chandrasekaran, J.; Brumin, M.; Wolf, D.; Leibman, D.; Klap, C.; Pearlsman, M.; Sherman, A.; Arazi, T.; Gal-On, A. Development of broad virus resistance in non-transgenic cucumber using CRISPR/Cas9 technology. Mol. Plant Pathol. 2016, 17, 1140–1153. [Google Scholar] [CrossRef] [Green Version]
  239. Wang, X.; Tu, M.; Wang, D.; Liu, J.; Li, Y.; Li, Z.; Wang, Y.; Wang, X. CRISPR/Cas9-mediated efficient targeted mutagenesis in grape in the first generation. Plant Biotechnol. J. 2018, 16, 844–855. [Google Scholar] [CrossRef]
  240. Santillán Martínez, M.I.; Bracuto, V.; Koseoglou, E.; Appiano, M.; Jacobsen, E.; Visser, R.G.F.; Wolters, A.A.; Bai, Y. CRISPR/Cas9-targeted mutagenesis of the tomato susceptibility gene PMR4 for resistance against powdery mildew. BMC Plant Biol. 2020, 19, 284. [Google Scholar] [CrossRef]
  241. Ortigosa, A.; Gimenez-Ibanez, S.; Leonhardt, N.; Solano, R. Design of a bacterial speck resistant tomato by CRISPR/Cas9-mediated editing of SlJAZ2. Plant Biotechnol. J. 2019, 17, 665–673. [Google Scholar] [CrossRef] [Green Version]
  242. 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]
  243. Khang, B. The regulatory status of genome-edited crops. Plant Biotechnol. J. 2016, 14, 510–518. [Google Scholar]
  244. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Aman, R.; Ali, Z.; Butt, H.; Mahas, A.; Aljedaani, F.; Khan, M.Z.; Ding, S.; Mahfouz, M. RNA virus interference via CRISPR/Cas13a system in plants. Genome Biol. 2018, 19, 1. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Global yield loss to pathogens and pests on major food crops.
Figure 1. Global yield loss to pathogens and pests on major food crops.
Ijms 23 14370 g001
Figure 2. Benefits of herbicide-resistant (a) and insecticidal (b) transgenic crops.
Figure 2. Benefits of herbicide-resistant (a) and insecticidal (b) transgenic crops.
Ijms 23 14370 g002
Table 1. Number of transgenic events of insect resistance authorized for cultivation and market.
Table 1. Number of transgenic events of insect resistance authorized for cultivation and market.
CropGene
Introduced
Doner
Species
Target InsectReleased Event
LepidopteranColeopteranHemipteran
CottonCryB. thuringiensis23 (1) 125
vip31
CpTIV. unguiculata(1)
CowpeaCryB. thuringiensis1 1
EggplantCryB. thuringiensis1 1
MaizeCryB. thuringiensis15 (2)8 (3) 25
vip3B. thuringiensis2 (1)
RNAi of Snf7D. virgifera (1)
PotatoCryB. thuringiensis 30 30
RiceCryB. thuringiensis3 (2) 3
SoybeanCryB. thuringiensis4 (2) 4
SugarcaneCryB. thuringiensis3 3
TomatoCryB. thuringiensis1 1
 Total5438193
Note: The brackets indicate the numbers of the genes introduced for pyramiding.
Table 2. Number of RNAi transgenic events of viral resistance authorized for cultivation and market.
Table 2. Number of RNAi transgenic events of viral resistance authorized for cultivation and market.
CropGene IntroducedTarget VirusReleased Event
BeanReplicase gene bgmv_ac1Bean golden mosaic virus1
PapayaCoat protein gene prsv_cpPapaya ringspot virus3
Replicase gene prsv_rep1
PlumCoat protein gene ppv_cpPlum pox virus1
PotatoCoat protein gene pvy_cpPotato virus Y8
Replicase gene plrv_orf1Potato leaf roll virus7
Helicase gene plrv_orf2(7)
SquashCoat protein gene zymv_cpZucchini yellow mosaic virus2
Coat protein gene cmv_cpCucumber mosaic virus[1]
Coat protein gene wmv_cpWatermelon mosaic virus 2[2]
Sweet pepperCoat protein gene cmv_cpCucumber mosaic virus1
TomatoCoat protein gene cmv_cpCucumber mosaic virus1
Total25
Note: The parentheses and square brackets indicate the numbers of the genes introduced for resistance pyramiding or broadening, respectively.
Table 3. CRISPR-Cas edited improvement of crops for disease resistance.
Table 3. CRISPR-Cas edited improvement of crops for disease resistance.
CropGene ModifiedResistance ImprovedReference
CitrusCsLOB1Citrus canker[236]
CassavanCBP-1, nCBP-2Cassava brown streak disease[237]
CucumbereIF4ECucumber vein yellowing virus[238]
GrapeVvWRKY52Botrytis cinerea[239]
RiceBsr-k1Broad spectrum resistance[234]
OsSWEET11, 13, 14Bacterial blight[235]
TomatoPmr4Powdery mildew[240]
Jaz2Bacterial speck disease[241]
WheatTaMlo-A1, -B1, -D1Powdery mildew[220]
TaEdr1 (three homologs)[242]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, H.; Wang, Y.; Fu, F.; Li, W. Transgenic Improvement for Biotic Resistance of Crops. Int. J. Mol. Sci. 2022, 23, 14370. https://doi.org/10.3390/ijms232214370

AMA Style

Yu H, Wang Y, Fu F, Li W. Transgenic Improvement for Biotic Resistance of Crops. International Journal of Molecular Sciences. 2022; 23(22):14370. https://doi.org/10.3390/ijms232214370

Chicago/Turabian Style

Yu, Haoqiang, Yingge Wang, Fengling Fu, and Wanchen Li. 2022. "Transgenic Improvement for Biotic Resistance of Crops" International Journal of Molecular Sciences 23, no. 22: 14370. https://doi.org/10.3390/ijms232214370

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