Next Article in Journal
Parallel Differentiation and Plastic Adjustment of Leaf Anatomy in Alpine Arabidopsis arenosa Ecotypes
Next Article in Special Issue
Gamma Radiation Induced In-Vitro Mutagenesis and Isolation of Mutants for Early Flowering and Phytomorphological Variations in Dendrobium ‘Emma White’
Previous Article in Journal
Biochemical Insights into the Ability of Lemna minor L. Extract to Counteract Copper Toxicity in Maize
Previous Article in Special Issue
Chili Pepper AN2 (CaAN2): A Visible Selection Marker for Nondestructive Monitoring of Transgenic Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genome Editing for Sustainable Crop Improvement and Mitigation of Biotic and Abiotic Stresses

1
Centre for Research in Biotechnology for Agriculture, Universiti Malaya, Kuala Lumpur 50603, Malaysia
2
Department of Cell and Molecular Biology, Faculty of Biotechnology and Biomolecular Sciences, Universiti Putra Malaysia, Serdang 43400, Malaysia
*
Author to whom correspondence should be addressed.
Plants 2022, 11(19), 2625; https://doi.org/10.3390/plants11192625
Submission received: 17 August 2022 / Revised: 22 September 2022 / Accepted: 30 September 2022 / Published: 6 October 2022
(This article belongs to the Special Issue Plant Biotechnology and Crop Improvement)

Abstract

:
Climate change poses a serious threat to global agricultural activity and food production. Plant genome editing technologies have been widely used to develop crop varieties with superior qualities or can tolerate adverse environmental conditions. Unlike conventional breeding techniques (e.g., selective breeding and mutation breeding), modern genome editing tools offer more targeted and specific alterations of the plant genome and could significantly speed up the progress of developing crops with desired traits, such as higher yield and/or stronger resilience to the changing environment. In this review, we discuss the current development and future applications of genome editing technologies in mitigating the impacts of biotic and abiotic stresses on agriculture. We focus specifically on the CRISPR/Cas system, which has been the center of attention in the last few years as a revolutionary genome-editing tool in various species. We also conducted a bibliographic analysis on CRISPR-related papers published from 2012 to 2021 (10 years) to identify trends and potential in the CRISPR/Cas-related plant research. In addition, this review article outlines the current shortcomings and challenges of employing genome editing technologies in agriculture with notes on future prospective. We believe combining conventional and more innovative technologies in agriculture would be the key to optimizing crop improvement beyond the limitations of traditional agricultural practices.

1. Introduction

Climate change, such as extreme weather or temperature, drought, increasing soil salinity, and flooding, significantly affects the food production system, posing serious threats to food security. The adverse effects of climate change on agricultural productivity have been reported in several regions, including Asia [1], sub-Saharan Africa [2], and the European Union (EU) [3]. For example, the heatwave and drought in the EU in 2018 have reduced cereal production by 8% compared to the previous five-year average [4], causing fodder shortages for livestock and increasing commodity prices. The impacts of climate change on agriculture in developing countries are more significant than in developed countries, mainly as these countries are located in tropical latitudes, which are more sensitive to climate change [5]. In addition, differences in vulnerability between these regions might be due to differences in endowments of human skills, physical infrastructure, and rapid demography growth, causing developing countries to have lower levels of resilience [6,7,8]. Ensuring sustainable crop production and food security has become challenging not only due to the growing environmental pressures but also the ever-increasing human population. Around 720 to 811 million people, about a tenth of the global population, still suffer from hunger. Meanwhile, more than 2 billion people are in the ‘food insecure’ category [9]. Another 130 million people may be added to the latter category due to the recent COVID-19 pandemic [10]. These problems will continue to worsen with the projected global population growth since the yield of grain crops, such as rice, wheat, and maize, has already reached a plateau [11]. With an estimated world population of 9.7 billion by 2050, crop productivity will need to increase by another ~70% while simultaneously reducing the environmental impacts [12]. Moreover, climate change increases the severity of biotic and biotic stresses on crops. Biotic stresses, such as pathogens, insect pests, and weeds, cause average output losses ranging from 17.2% in potatoes to 30.0% in rice [13]. Likewise, abiotic stresses, such as temperature extremes, drought, and lack of nutrient deficiency, caused the loss of 51–82% of the global crop output annually [14]. As the intensity of biotic and abiotic stresses on crops increases because of climate change, novel approaches are required to enhance plant tolerance. Given that the conventional agricultural practices are inadequate to meet current and future food demands and deal with the aggravated impacts of biotic and abiotic stresses due to climate change, developing practical and effective adaptation strategies is indispensable to enhance crop productivity and ensure food security. Ideally, the strategies driving this effort should be sustainable and environmentally friendly while minimizing adverse environmental impacts.
Crop breeding, including cross-breeding and mutation breeding, has been used to enhance crop performance under climate change scenarios. However, breeding programs can be laborious and time-consuming, even aided by marker-assisted selection. It can take 8 to 10 years [15] or 6 to 15 years [16] to produce a genetically superior cultivar for agricultural production. Plant breeders have used cross-breeding based on naturally occurring mutations [15] or mutation breeding techniques based on ionizing radiation and chemical mutagens to generate new varieties with desired agronomic traits, including improved stress-tolerance potential and biofortification [17]. Nevertheless, since cross-breeding is limited to traits present in the parental genomes, low variability in elite germplasms restrains the use of this technique. The outcomes of the mutation breeding technique are unpredictable even though lower mutation rates have been reported in essential genes compared to non-essential genes [18]. In addition, complex and tedious screening and selection procedures are required to identify the desired trait from a large population of mutagenized plants [19]. Transgenic technologies that involve transferring desired trait-coding genes into the elite cultivars are undoubtedly an alternative to counter losses in crop yield [20]. However, the time and expenses for developing a genetically modified (GM) crop with desirable traits are enormous. The major limitation of this method is the low public acceptance of GM crops and, related to this, the complex and strict safety regulatory procedures [21]. In addition, different countries have adopted different regulatory procedures. However, to date, only a few countries, such as Switzerland, strictly restricted or legally prohibited the cultivation of GMOs [22].
Given the importance of securing sustainable crop yield, the challenge now is to improve the existing technologies or develop alternative technologies/solutions to increase crop yields. Here, we discuss the possibility of using genome editing, particularly the CRISPR/Cas9 system, to alleviate the impact of environmental stress and enhance crop production. A bibliometric analysis of CRISPR-related articles published in the SCOPUS database was done to evaluate its current trend of publications from 2012 to 2021. The selected timeline represents the first decade since the discovery of CRISPR/Cas9 in 2012 for use in genome editing [23]. This content analysis allows us to identify certain ‘hot spots’ or themes and reveal the potential of CRISPR-related research in plants.

2. Genome Editing Technologies

Genome editing techniques using sequence-specific nucleases (SSNs) have become popular in plant research. They have been used to develop high-yielding crops, improve the adaptability of crops to environmental stresses or enhance their nutrition content [24]. To date, there are four SSNs, namely meganucleases, zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced palindromic repeat (CRISPR)/CRISPR-associated (Cas) protein systems (Figure 1). These technologies allow precise targeting and modifying of specific DNA sequences in three common steps: (1) an exogenous engineered nuclease consisting of a recognition module and nuclease domain recognizes the target DNA sequence, (2) the engineered nuclease binds to the target DNA sequence and induces double-strand breaks (DSBs) at or in the vicinity of the target site and (3) the DSBs will then be repaired by either non-homologous end-joining (NHEJ) or homologous recombination (HR). NHEJ is an error-prone repair mechanism that often results in insertion and deletion (Indel) mutations, whereas HR results in a precise repair of DSBs [25].
Meganucleases were the first SSN used to create targeted DSBs in eukaryote genomes [26]. They are naturally occurring endonucleases found in prokaryotes, archaea, and unicellular eukaryotes [27]. The first meganuclease, I-SecI, was discovered in yeast [28]. Meganucleases are the most specific naturally occurring endonucleases as they recognize 14–40 bp long DNA sequences [29]. These enzymes have a larger recognition site than the type II restriction enzymes used in recombinant DNA technology. Due to their long recognition sequence and high specificity, meganucleases can efficiently target and modify any sequence of interest [30]. For these reasons, meganucleases have been used to create targeted DSBs in eukaryote genomes since the 1990s. In 1993, Puchta and colleagues published a landmark paper demonstrating that I-SecI-induced DSBs enhance HR in Nicotiana tabacum [31]. This marked the arrival of precise genetic engineering in plants using SSNs. Since then, several efforts have been made to introduce trait genes into plants. For example, D’Halluin et al. [32] inserted multiple trait genes, 4-hydroxyphenylpyruvate dioxygenase (hppd) and modified enol-pyruvylshikimate-3-phosphate synthase genes (epsps) into cotton using meganucleases to enhance its herbicide tolerance. Although meganucleases have been successfully applied for targeted gene editing in the plant, they have a few limitations, such as the low catalytic activity of the enzyme, prone to sequence degeneracy, and the lack of mature DNA binding structure, which hinders their wide applications [33,34].
ZFNs are fusions of the DNA recognition domain of zinc finger protein and the cleavage domain of the FokI endonuclease [35]. ZFNs act through DNA/protein recognition, and each zinc finger recognizes three base pairs (bp). As FokI must dimerize to become active, ZFNs should be designed as a pair to ensure the correct orientation and appropriate spacing for FokI dimerization [36]. To date, ZFN-mediated gene modification has been reported in various crops, such as soybean [37], maize [38], wheat [39], and rice [40]. Yet, their application as an editing tool in crops is limited because of the complexity and cost of the protein construction for each targeted site, and the potential cytotoxicity effects, presumably due to cleavage at off-target sites [41].
Similar to ZFNs, TALENs comprise transcriptional activator-like effector (TALE) repeats (comprise the DNA binding domain) and a FokI endonuclease (comprises the cleavage domain) [30]. TALEs are type III effector proteins derived from Xanthomonas spp. Their DNA binding ability was first reported in plants in 2007 [42,43]. In 2009, the recognition code of TALE targeting DNA sequence was also decrypted [44]. The DNA binding domain in TALE monomers contains a central repeat domain, which consists of tandem repeats of 34 amino acid residues. Each 34-amino-acid-long repeat in the central repeat domain targets only one nucleotide in the target DNA sequence. This made TALENs a better gene-editing tool compared to ZFNs as they allow flexible target design. Two hypervariable amino acid residues at the 12th and 13th positions are highly variable (termed as repeat variable di-residue [RVD]) and critical for specific nucleotide recognition.
TALENs have been demonstrated in various plant species, such as Arabidopsis, tobacco, soybean, sugarcane, maize, and wheat [45,46]. The use of TALENs in crop improvement was first reported in rice, where OsSWEET14 (bacterial blight susceptibility gene) was disrupted, and the resulting mutant rice displayed bacterial blight resistance [47]. Other applications of TALENs in crop improvement include producing flavor in rice [48], developing powdery mildew resistant wheat [49], enhancing the nutrient content of soybean [50], and increasing anthocyanin levels in tomatoes [51]. However, despite their potential for crop improvement, several challenges of TALENs have limited their applications. One major drawback is the inefficient delivery of the TALEN system into a cell due to the large size of cDNA encoding TALEN (about 3 kb). Furthermore, the construction of TALE repeats remains a bottleneck and the efficiency of TALENs targeting a gene is variable [45].
The discovery of the CRISPR/Cas9 genome editing system has revolutionized the fields of functional genomics in animal and plant biology. Originating from bacteria and archaea as an adaptive immunity system, the CRISPR/Cas9 system has become a viable tool for targeted genome editing in prokaryotes and eukaryotes.

3. The CRISPR/Cas System

In 1987, CRISPR was discovered accidentally in the Escherichia coli genome while Ishino et al. [52] were sequencing the iap gene encoding alkaline phosphatase isozyme conversion enzyme. Downstream the iap gene, the authors discovered a unique set of tandemly repeating 29-nucleotide (nt) DNA sequences interspersed with 32-nt spacer sequences. They were unaware of the biological role of these repeats due to the lack of sequence homology with other known sequences. Later in 1993, long tandem repeats were discovered by Mojica et al. while sequencing several Haloferax mediterranei genome segments [53]. This marks the first finding of archaeal direct repeats. The series of interspaced repetitions was later classified as clustered regularly interspaced short palindromic repeats (CRISPR) [54]. As biological science advances towards the genomic era, CRISPR has now been identified in various bacterial and archaeal genomes.
In 2005, it was finally revealed that spacers present within CRISPRs were produced by invading phages and plasmids [55,56]. This established the CRISPR/Cas system as an adaptive immune system of bacteria and archaea that defends bacteria from bacteriophages and mobile genetic elements (MGEs) by eliminating invasive genomic elements [57]. CRISPRs can prompt the capture of invading DNA fragments to serve as a record of prior genetic aggressions [55,56]. The significance of CRISPR/Cas systems as adaptive immunity has been reinforced by subsequent experimental findings, which reiterated that new spacer sequences from the infecting phage are acquired by the bacterial CRISPR array [58,59,60,61].
The CRISPR/Cas systems were initially classified into three types (Types I, II and III) based on proteins and accessory RNAs. Type I and III systems use a complex of multi-Cas protein for target DNA recognition and cleavage, whereas the Type II system relies on a single Cas9 protein to accomplish the interference [62]. Further experimentation and analysis have further divided the classification into 2 classes, 6 types and 33 subtypes [63]. As the classification of CRISPR has been described in earlier reviews [64,65], we are not explaining it in detail in this paper.
The type II CRISPR/Cas system is the most widely used and best studied due to its straightforward constructs compared to the other systems. Type II CRISPR/Cas system employs a single protein, Cas9, and two non-coding RNAs, CRISPR RNA (crRNA) and trans-activating crRNA (tracrRNA), for target recognition and cleavage. The dualtracrRNA:crRNA guides the Cas9 nuclease to recognize protospacer adjacent motifs (PAMs or 5′-NGG-3′) on the target DNA sequence. Cleavage of the target DNA is then performed by two Cas9 nuclease domains, the HNH domain (cleave the DNA strands complementary) and RuvC-like domain (cleave the non-complementary). Both induce a DSB three bases upstream of the PAM site of the target region.
The newly established CRISPR/Cas system replaces the dual-tracrRNA:crRNA with a single guide RNA (sgRNA) to ease the genome modifications. With this system, one can perform genome editing by simply modifying the 20-nucleotide sgRNA to be complementary to the target DNA. Overall, a CRISPR/Cas9 project involves the steps below:
(i) Target and PAM sequence identification;
(ii) Evaluate off-target effects;
(iii) sgRNA synthesis;
(iv) Cloning of the sgRNA into a suitable plant expression plasmid;
(v) Plant transformation and screening of the edited lines.
As described above, CRISPR-mediated genome editing involves the generation of Cas9-induced DSBs that are repaired by NHEJ or HR. For developing agronomic traits, precise genome modification is required. Although HR can be used to precisely repair the DSBs when DNA donor templates are supplied, it is rarely used in crop improvement because of its low efficiency in higher plants. The recent developed powerful technologies, i.e., base editing and prime editing, have partly overcome such barriers and greatly improved crop breeding opportunities.
Base editing (BE) is a technique that directly converts one target genomic DNA base into another at a targeted locus without producing a double-stranded break. Combining cytosine or adenine deaminases with CRISPR-Cas9, a range of cytosine base editors (CBEs) and adenine base editors (ABEs) has been developed in recent years [66,67,68]. These varieties allow exact C-to-T or A-to-G base conversions without causing a DSB. Using CBE, cytosine (C) is deaminated to create uracil (U). The uracil (U) is read as thymine during DNA replication (T). CBE consequently provides a single-base substitution from CG to TA [69]. In ABE, the inactive CRISPR–Cas9 domain is connected to adenosine deaminase, which helps convert adenine (A) to inosine, unlike cytidine deaminase in CBE. This inosine is read during DNA replication as guanine (G). Consequently, ABE generates AT to GC base substitutions [70]. Since their discovery, base editors have become valuable tools for precisely modifying the genomes of eukaryotic organisms [71,72,73,74].
Prime editing (PE) is another innovation made to the genome editing toolbox. Previous BE procedures created single base substitutions for four transitions (C > T; T > C; A > G; G > A), and newer studies included two transversions (C > G and G > C). Instead of a deaminase, PE uses an extended guide RNA (pegRNA)-guided reverse transcriptase, which allows pegRNA to install substitutions, insertions, and deletions [75]. In contrast, PE contains all 12 alterations, including the eight transversions. This increases the versatility and robustness of the gene editing strategy. Although still in its infancy, PE applications show promise in multiple cell types, organoids, and mice embryos. Data on its application in plant systems have also started to emerge. In maize, PE has introduced W542L and S621I double mutations in two ALS genes, ZmALS1 and ZmALS2, which may confer resistance to several ALS-inhibiting herbicides. In rice, the level of PE efficiencies ranged from 2.22 to 31.3% [76]. In one experiment, triple amino acid substitutions (T169I, A170V, and P173S) were introduced into OsEPSPS [77], which may confer a higher level of glyphosate resistance [78].
In addition to the Cas9 protein, three more family proteins, namely Cas12, Cas13, and Cas14, as well as their orthologs, have been identified. Cas12 family protein is considered more advanced and versatile than Cas9 due to several characteristics, such as smaller size, lack of need for trans-activating crRNA (tracrRNA), and ability to cleave DNA via its RuvC domain. In addition, it can edit many genes from a single RNA transcript due to an intrinsic RNAse that can process its own guide RNA array [79,80,81]. Type IV Cas13 has an RNA-guided RNase domain that could degrade nearby single-stranded-RNA (ssRNA) molecules [82]. It has been applied for targeted RNA interference in various organisms, including animals, human viruses, and plants [83]. Cas14 is exceptional for sequence detection since it does not need a PAM and is particularly sensitive to mismatches in the center of its target region [84].
Given the simplicity, versatility, and efficacy of the CRISPR/Cas9 system, this technology shows great potential for target mutagenesis in various plant species. Despite these advantages and significant developments in the CRISPR/Cas system, continuous efforts to improve its efficiency and practicality in agriculture are still desirable. Figure 2 summarizes the general procedures involved in plant transformation and CRISPR-based plant genome editing.

4. CRISPR/Cas9 for Genome Editing in Crops

The CRISPR/Cas9 system has been used in various crops to develop desirable and heritable traits, such as yield improvement, and biotic and abiotic stress management. Table 1 summarizes the applications of the CRISPR/Cas9 for crop improvement.

4.1. Abiotic Stress

Climate change leads to various abiotic stresses, threatening agricultural food production worldwide [101]. About 90% of all arable lands are prone to single or multiple abiotic stresses, such as water stress, extreme temperature, and salinity [102]. To survive, plants have evolved various mechanisms to respond to and cope with these stresses [103]. However, the plant stress-responsive and adaptation mechanisms are complex and governed by various genes, posing challenges to developing novel cultivars using the conventional methods [104]. As such, targeted genome editing on a single or multiple target sites through the CRISPR/Cas9 system could be a promising approach to developing abiotic stress-resilient crop varieties [25].
The CRISPR/Cas9 approach has been exploited to improve crop survival under adverse environmental stresses. For example, Zhang et al. [105] developed salinity-resistant rice through the CRISPR/Cas9 approach. By knocking out the OsRR22 gene, the authors found that the generated rice showed better plant growth than wild-type under salinity conditions [105]. A recent study indicated that OsNAC041 is a critical transcription factor involved in the salt stress response in rice. A targeted osnac041 mutant obtained using the CRISPR/Cas9 method showed a higher plant height than wild-type [106]. Other studies demonstrated that members of the AP2/ERF domain containing the RAV (related to ABI3/VP1) transcription factor family are involved in salinity stress adaption [107,108]. For instance, when the rice was exposed to salt stress, the OsRAV2 gene was activated. To determine the role of the GT-1 element in the OsRAV2 gene, Duan et al. [109] designed a sgRNA targeting the GT-1 region of the promoter. They found that the mutant lines could not express the OsRAV2 gene under salinity conditions, confirming the importance of this gene in response to salinity stress. A similar finding has been reported by Liu et al. [110], where the CRISPR/Cas9-mediated OsGTγ-2 knockout lines showed salt-hypersensitive phenotypes. Besides rice, the CRISPR/Cas9 genome editing technology has also been applied to other crops, such as wheat [111], soybean [112], maize [113], and tomato [114].
Drought stress disturbs physiological and biochemical processes in plants, limiting plant growth and yield [115]. Several genes and phytohormone signaling pathways have been shown to play critical roles in drought stress responses. Of these, abscisic acid (ABA) is a central regulator of water use and coordinates the plant’s responses to drought stress. Hence, several studies have been conducted to improve drought tolerance in crops by targeting the genes involved in ABA signaling. For example, Zhang et al. [116] determined the role of OsABA8ox2, which encodes ABA 8′-hydroxylase, in rice drought tolerance. The authors found that the CRISPR/Cas9-mediated OsABA8ox2 knockout lines showed increased drought-induced ABA in roots and induced root formation beneficial to drought tolerance. In contrast, overexpressing OsABA8ox2 in rice suppressed root elongation and exhibited hypersensitivity to drought stress [116]. The ENHANCED RESPONSE TO ABA1 (ERA1), which encodes the β-subunit of the protein farnesyltransferase, was mutated in Japonica rice cv. Nipponbare using the CRISPR/Cas9 system [117]. The rice osera1 mutant lines showed increased sensitivity to ABA and drought tolerance through stomatal regulation, suggesting that ERA1 could be a potential candidate gene for enhancing drought tolerance in crops. Another study by Yin et al. [118] showed that the OsEPFL9 (Epidermal Patterning Factor like-9) mutants had more than an eight-fold reduction in stomatal density (SD) in the CRISPR/Cas9-edited rice plants. The reduced SD allows the edited rice lines to resist drought stress. Under optimal conditions, a significant reduction in carbon assimilation and conductance and enhanced water use efficiency (WUE) was observed when SD was reduced by 50% in barley and wheat [119,120]. Likewise, in well-watered conditions, a CRISPR-based knockout of grapevine VvEPFL9-1 reduced SD by 60% and caused reduced carbon assimilation as compared to WT [121]. In tomatoes, slmapk3 mutants generated through CRISPR/Cas9 showed that SlMAPK3 is involved in drought response, and the slmapk3 mutants showed more severe wilting symptoms and suffered cell membrane damage under drought stress [122].
Some studies used the CRISPR/Cas9 technology to reduce mineral toxicity. For example, Nieves-Cordones et al. [123] developed low cesium-containing rice plants by inactivating the K+ transporter OsHAK1 using the CRISPR/Cas9 system. In rice, knocking out OsARM1 and OsNramp5 showed improved arsenic tolerance [124] and low cadmium accumulation [125], respectively. Another example of increasing plant stress resistance was shown by Shao et al. [126], where the authors developed a semi-dwarf variety of bananas using the CRISPR/Cas9 system to disrupt the genes responsible for the gibberellin biosynthesis. As a result, the developed bananas are more resistant to storms and heavy wind. Besides generating knockouts on the susceptible genes, genome-editing tools can also be used for knock-ins of a desirable gene. For instance, Shimatani et al. [127] used CRISPR/Cas9 to insert a maize promoter before the drought tolerance gene, ARGOS8. Consequently, the edited maize crops showed a greater grain yield during water stress.
These studies demonstrated that the CRISPR/Cas system could edit the plant genome, allowing us to investigate the role of genes involved in response to abiotic stresses. However, reports on targeting abiotic stress tolerance genes are scarce, primarily due to the complexity associated with abiotic stress tolerance, often involving the modulation of several genes to alter the trait of interest.

4.2. Biotic Stress

Plants are constantly plagued by pathogens, such as viruses, bacteria, and fungi, which can significantly reduce crop quality and yield [128]. The majority of disease-resistant crops against non-viral diseases are produced through genome editing and targeted mutagenesis of genes that negatively influence defense [129]. While few such genes are available for increasing disease resistance, many of these loci have already been successfully exploited for increased resistance.
In rice, genome editing has shown a remarkable result in combating diseases using CRISPR/Cas9. Most pathogens use the sucrose transporters that are encoded by the SWEET gene family in many plants [130]. In two experiments, CRISPR/Cas9 was utilized to target the promoter region of a few OsSWEET genes to develop resistance against bacterial leaf blight [131,132]. Knockout of the OsERF922 gene that expresses ethylene response in the plant using CRISPR/Cas9 reduced the effect of leaf blast disease, thereby enhancing its tolerance toward it [133]. Additionally, CRISPR/Cas9 editing of the eukaryotic elongation factor, eIF4G, in rice resulted in plants that were immune to the rice tungro virus [134]. The infected CRISPR-edited plants contained no detectable viral proteins and produced better yields than wild-type plants.
The advancement of the CRISPR/Cas9 system has furthered the development of resistance to multiple diseases at the same time. Engineering the broad spectrum of disease resistance in staple crops on a large scale could provide a single solution to several diseases that are affecting crop production [131]. The editing of bsr-k1, a rice gene that binds to and increases the turnover of defense-related genes [135], is an example of this strategy. By “turning off” these critical defense genes, edited rice plants were resistant to both leaf blast and bacterial leaf blight. When challenged with rice leaf blast in the field, the transgenic lines show a greater yield of 50% more without affecting other agronomic features [135]. Likewise, the same strategy has also been applied to other crops for disease resistance. For example, broad-spectrum resistance was obtained by altering a single locus in tomatoes [136]. The SlDMR6-1 mutations by CRISPR/Cas9 in the edited lines maintain an increased salicylic acid level in the plant with a significant reduction of disease symptoms and pathogen abundance, gaining resistance to Pseudomonas syringae, Phytophthora capsici, and Xanthomonas spp. [136]. In barley, CRISPR/Cas9-mediated editing of MORC1, a defense-related gene, increased resistance to barley powdery mildew and Fusarium graminearum [137]. In addition, the authors showed that the edited barley plants had lower levels of fungal DNA and fewer lesions.
In some species, targeting homologs of Mildew-resistance Locus (MLO) and other loci enhanced the resistance to these fungal infections. By concurrently targeting the three homologs of the MLO, TaMLO-A, TaMLO-B, and TaMLO-D, CRISPR/Cas9 can increase the resistance of wheat to powdery mildew [49]. Another example is the Tomelo transgene-free tomato, which is resistant to powdery mildew disease and was produced by targeting SlMlo1 gene using CRISPR/Cas9 [138]. Zhang et al. [139] changed the three homologs of the wheat TaEDR1 gene simultaneously using CRISPR/Cas9 to improve resistance to powdery mildew disease. In grapevine, targeting of the MLO homologs boosted the resistance to powdery mildew, whereas the edited line of grapevine had a two-fold reduction in powdery mildew sporulation [140]. In other efforts, knockout of the 14-3-3 c and 14-3-3 d protein simultaneously, a negative regulator of disease response, in cotton enhanced resistance to Verticillium dahliae [141]. The edited cotton showed fewer disease symptoms and lowered pathogen presence compared to control [141].

4.3. Yield

One of the essential keys to sustaining food production is crop yield. It is the most direct means to address the ever-rising food demand from a growing population. However, crop yield is a complex trait regulated by many factors. Therefore, much research has been done to identify the quantitative trait loci (QTLs) associated with morpho-agronomic and yield-related traits in various crop plants [142].
One way genome editing can increase crop yield is to eliminate genes that have a detrimental impact on yields, such as genes limiting grain size and weight [143,144]. In one recent example, CRISPR/Cas9 was used to individually knock out the genes of four negative yield regulators (Gn1a, DEP1, GS3, and IPA1) in the rice cultivar Zhonghua 11. Each of the individual knockout mutants, Gn1a, DEP1, and GS3, showed increased yield characteristics in the T2 generation [145]. Similarly, Xu et al. [146] used a CRISPR/Cas9-mediated multiplex genome-editing technology to knock out three main rice negative regulators of grain weight (GW2, GW5, and TGW6) simultaneously, and the resulting mutants had a considerable increase in thousands of grain weights. In another study on wheat, CRISPR/Cas was used to knock out the three homoeoalleles of GASR7, and the mutant plant produced a much heavier kernel weight when compared to the wild-type wheat plants [147]. Besides grain, targeting a tomato cis-regulatory region in the CLAVATA-WUSCHEL stem cell circuit (CLV-WUS) using CRISPR/Cas9 resulted in an edited tomato with an increased number of locules (seed compartments) and bigger fruit size [148].
Alternatively, genome editing can also influence crop yield through other strategies. CRISPR/Cas9 technology was employed in maize to create high amylopectin variants from superior cultivars by knocking out the waxy gene [149]. The edited maize cultivars yielded 5.5 bushels per acre more than conventionally bred high amylopectin varieties. Furthermore, they could be developed in a shorter time, demonstrating the feasibility of genome editing in particular specialized applications [149]. Furthermore, reducing the ABA response of rice plants can also enhance the yield. Rice plants with simultaneous mutations of class I PYL genes (encoding receptors for ABA) using CRISPR/Cas9 had better yields than the control [95]. Under well-watered conditions, triple knockout of PYLs 1,4,6 resulted in a 30% increase in yield [95]. It is interesting to see how these ABA-encoding PYL genes affect yield under less-optimal conditions. A recent study shows that under drought conditions, the wheat PYL1-1B (TaPYL1-1B) is responsible for increased yield and drought resistance, where it exhibited higher ABA sensitivity, photosynthetic capacity and WUE [150].
A higher yield of tomatoes can also be achieved by modifying the flower repressor gene using CRISPR/Cas9. Knockout of the flowering repressor SELF-PRUNING 5G (SP5G) gene produced tomato plants that have rapid flowering, which in turn yield earlier with compact determined growth [151]. In contrast, mutations in the SELF PRUNING (SP) gene changed the plant architecture to a bushier state with more branches [152]. The resultant mutants with two modifications had faster flowering time and earlier fruit ripening than the control lines. In another study, CRISPR-based knockout of tomato SlAGL6 enhanced yield under heat stress. The tomato agl6 mutants displayed facultative parthenocarpy without any pleiotropic effect and produced seedless fruits of equal weight and shape to WT [153]. Under salinity stress, the CRISPR-edited soybean gmaitr mutants yield was much less affected than the WT in plant height, number of pods per plant, and seed weight [112]. The number of studies on plant yield and resilience improvement is expected to grow, in line with the rapid advancement of genome editing tools.

5. VOSviewer Bibliometric Analysis

We used ‘Visualization of similarities (VOS) viewer’ version 1.6.17; [154] to visualize and analyze the bibliometric network of CRISPR-related publications extracted from the SCOPUS database for the past 10 years (2012 to 2021). VOSviewer is a handy tool that allows a graphical representation and interpretation of networks representing co-authorship, journals, institutions, or co-occurring keywords based on a selected topic of interest [155].
Based on our keyword search, more than 5200 scientific papers focused on “CRISPR” OR “genome editing” AND “plants” have been published in the last ten years (2012–2021). We compiled a list of relevant publications with the co-occurrence of keywords in the title, abstract, and keywords sections from all publication types (2012–2021), including journal articles, books, and conference proceedings, to generate bibliographic maps and networks using the software VOSviewer. The criteria were set as follows: the keywords repeated at least five times were selected, singular and plural forms were standardized to singular forms to avoid redundancies, and full names and abbreviations were standardized to full names. Interchangeable keywords (e.g., ’corn’ and maize’) and spelling differences (e.g., ‘colour’ vs ‘color’) were also standardized in the ‘thesaurus’ option before running the bibliographic analysis. Based on these premises, 50 keywords were used and clustered according to their strength of association. Four clusters (sets of closely related nodes) were generated and integrated into a network overlay visualization map. The maps and networks for the analyses are presented in Figure 3. The list of 50 keywords based on their ranking is presented in Table 2.
Table 2 shows that “crispr”, “CRISPR/Cas9”, and “plant protein” are the three most used keywords in CRISPR-related plant research publications. Of several different types of plants/crops (e.g., model plants, food crops, industrial crops, and ornamental plants) [156], only the model plants (Arabidopsis and tobacco) and food crops (rice, tomato, wheat, maize and soybean) are present in the network map. Their total number of occurrences (shown in brackets) are as follows: Arabidopsis (673), tobacco (192), rice (525), tomato (239), wheat (224), maize (213), and soybean (134).
Four CRISPR system-related keywords, “crispr” (2386), “cas” (70), “crispr/cas” (144), and “CRISPR/Cas9” (821), have been identified in the top 50 keywords during the keyword search. These keywords were not grouped in the ‘thesaurus’ option before the analysis since each keyword may represent a unique value. The highest cited keyword, “crispr” (2386 occurrence), may represent the investigation of CRISPR as a biological phenomenon in the bacterial immune system, which later formed the basis of “CRISPR/Cas” technology. After the discovery of the CRISPR/Cas technology, studies on CRISPR as a biological phenomenon have continued to provide knowledge to further improve the CRISPR/Cas system applications.
Further advancements in the CRISPR/Cas systems are oriented to expand its applications to other organisms and cell types and identify other alternatives to Cas9 proteins to improve CRISPR editing scope and specificity [156]. This is reflected by the presence of the keyword “cas” in the network map. The three most distinguishable variants of Cas proteins identified so far are Cas3 in type I systems, Cas9 in type II systems, and Cas10 in type III systems [157]. In addition, many other Cas proteins have also been harnessed to expand the CRISPR/Cas targeting scope, including Cas9, Cas12a, and Cas13 variants and orthologs. The expanding list of these Cas proteins and their applications has been covered extensively in recent reviews [64,158,159,160,161].
The closeness of the keyword “CRISPR/Cas9” with its surrounding keywords, such as “chromosome”, “gene”, “transgene”, “site-directed mutagenesis”, and “crop” indicated the application of the CRISPR/Cas system for the past decade as genome editing tools in crops, allowing specific and targeted changes in the gene of interest. A key approach in plant engineering for the past few decades has involved the integration of specifically assembled DNA cassettes or foreign genes into the host plant. The ability to express non-native segments of DNA in certain plants or crops resulted in novel plants with desirable traits such as herbicide resistance, pest resistance, and disease resistance [162]. It is also worth noting that transgene-free plants produced by genome editing using the CRISPR/Cas-based system (e.g., site-directed-nuclease-1 (SDN-1) type) is rapidly becoming its main selling point to avoid unnecessary regulatory issues and to gain better public perception [163,164]. These two factors are important for genome editing technology to be fully utilized and positively impact on the agricultural sector [165]. This may explain the relatedness of the keywords “crispr/cas” and “transgene” in the network map.
Genome editing reagent delivery into the host genome is a crucial topic in plant genome editing. The two most common ways of transferring a gene of interest into a host plant involved Agrobacterium-mediated transformation (AMT) or direct DNA transfer [166]. Both techniques aim to express the integrated transgene, silencing endogenous gene expression or modifying endogenous gene activity or function [167]. Compared to direct DNA transfer, AMT is more cost-effective and accessible to most plant researchers due to its low input (requiring low copies of DNA fragments carrying the gene of interest) and high throughput (high transformation efficiency) [168]. AMT also enables the transfer of large DNA fragments with minimal rearrangement, unlike the direct DNA transfer technique. These qualities made AMT the favored approach for plant transformation. This scenario is reflected by the network map (Figure 3), where the keyword “agrobacterium” stays in proximity to keywords, such as “crispr/cas”, “transgene”, and “crops”.
In contrast, keywords that may be related to physically or chemically direct DNA transfer methods, such as “biolistic delivery”, “gene gun”, “plant bombardment”, “electroporation”, “microinjection”, or “Polyethylene glycol (PEG)-mediated transfer” are not present in the network map. Another topic commonly associated with genome editing and CRISPR tools is using protoplasts (plant cells without cell walls) as a rapid validation system. It provides a platform to test the mutagenesis efficiency of RNA-guided endonucleases, promoters, sgRNA designs, or Cas proteins before the full-scale transformation in the host plant [169]. The popularity of this approach is reflected in the network map in which the keyword “protoplast” is present near “site-directed mutagenesis”, “crispr/cas”, and “transgene” nodes.
Apart from identifying the research “hot spots”, the network map in Figure 3 also revealed gaps in the current state of CRISPR research. For example, the lack of connecting lines and the relatively large distance between “CRISPR” and both plant organelles, “chloroplast” and “mitochondria” indicate the lack of CRISPR application in those organelles, as compared to its already wide application in the nuclear genome in various species. This scenario is probably due to the impermeability of those plant organelles to most RNA and DNA [170]. In addition, the delivery system of CRISPR reagents into the host plant genome remains a challenge in plant transformation. For example, the use of plant bombardment to deliver CRISPR/Cas components may not require a binary vector. However, this technique has other disadvantages, such as random integration into the plant genome, less editing efficiency, and being costlier compared to AMT [171].
Given that one of the main aims of plant genome editing is to mitigate the effects of climate change, it is quite surprising to see the absence of keywords related to environmental stresses (e.g., drought, extreme weather, and elevated temperature or carbon dioxide level). Perhaps these keywords are more used in the other sections (e.g., Introduction or Conclusion sections) and less frequently elaborated in detail in the sections extracted for this analysis (i.e., Title, Abstract, and Keyword sections). In summary, it is possibly safe to assume that the first decade of CRISPR/Cas research may have focused on the ‘foundation’ of the CRISPR/Cas editing system by making various technical improvements in its applications (e.g., the discovery of different Cas proteins, improvements on the delivery system, and evaluation of altered DNA and possible off-target mutations).

6. Limitations and Challenges

Genome editing technologies have been employed to make precise changes in plant genomes. They significantly impact both fundamental research and agricultural improvement [172]. Recent modification approaches, particularly CRISPR/Cas, have increased the effectiveness and feasibility of genome editing without the need for incorporating foreign DNA. However, there are still significant obstacles to these technologies in terms of efficient and practical application in crop improvement. One prominent limitation is the off-target effect of these technologies that rely on using SSNs for targeted disruption, insertion, or replacement of selected loci [173]. While most off-targets are caused by identical sequences homologous to the targeted sequence, these effects can also occur in the region close to the target site with unrelated sequences [174]. Efforts have been made to reduce the off-target effects, especially in the CRISPR/Cas system. For instance, many different alternatives of the traditional Cas9 protein have been introduced and developed for higher efficiency and lower off-target effects [175,176,177,178,179,180]. Others, such as base editors that allow for exact nucleotide modifications, epigenome modifiers that change DNA confirmation and related expression levels, and prime editing for precision insertion of small DNA segments, are all prospective options [181,182,183].
Another major challenge of utilizing genome editing technologies to develop improved crops is the stringent regulatory frameworks and extensive risk assessment procedures on GM crops [184]. Most nations have biosafety frameworks in place to govern GM crops generated using recombinant DNA technology. These biosafety frameworks often draw on the fundamental concepts for food safety and the environmental risk assessment of conventional GM crops inserted with foreign gene(s) with desired characteristic(s) [185]. However, with the advent of the gene-edited crop, it is necessary to reassess the present definition of GM crops and the accompanying regulatory frameworks, because different gene editing techniques may introduce different types of alterations in the plant genome. For example, SDN-3 mutation is more similar to the conventional recombinant technique, which introduces a whole transgene into the plant genome, therefore the final product is usually considered a GMO in many nations. In contrast, the SDN-1 can introduce single base substitutions, and in certain cases without the need for introducing DSB. As a result, the genetic features of certain gene-edited crops may differ from conventional GM crops, therefore requiring a case-by-case approach to assess the risk associated with each individual product of the genome editing event [186].
Currently, there is no common regulatory approach at the international level for genome editing because of the continuous debate over the similarities, and differences between gene-edited crops and conventional GM crops. Hence, many countries do not have a clear regulatory policy on the gene-edited crops produced, which further impedes the development and implementation of these improved crops in the field. Nevertheless, the broad use of gene-editing technology poses major technological problems for regulatory bodies to identify and distinguish the regulated crop, as it can be difficult to distinguish the naturally occurring edited events in the plant genome from artificial means. Therefore, an agenda supported by various entities such as experts, associations, regulators, and researchers are needed to address these complex issues and concerns raised by gene-editing in the plant for the benefit of all [187].

7. Conclusions and Future Prospects

Genome editing technologies can potentially improve plant agriculture and food production to feed the world’s growing population. Due to their efficiency, ease of engineering, and robustness, CRISPR/Cas systems have revolutionized plant genome engineering and globalized its applications. The current consensus is that CRISPR/Cas systems have the potential to improve plants and crops in various ways, such as increasing crop quality and yield, introducing abiotic stress-resistant traits (such as drought-, herbicide-, and insecticide-resistance), improving food safety by removing the need for an antibiotic-resistant marker, and prolonging food product shelf-life.
The main findings from the bibliographic analysis can be summarized as follows: (1) CRISPR/Cas systems are mainly used for nuclear genome editing. In addition to the nuclear genome, further development and applications of the CRISPR/Cas tools in the plant organelles (i.e., mitochondria and chloroplast genomes) are expected to increase as the technology advances, (2) most CRISPR/Cas editing so far is done on model plants (e.g., Arabidopsis and tobacco) or food crops (e.g., rice, tomato, and wheat). Discovery of novel Cas variants and orthologs and other CRISPR reagents should further expand the targeting scope of CRISPR/Cas systems to other types of plants/crops irrespective of species, (3) the issue of ‘transgene’ usage is one of the most widely discussed in the field of genome editing. Emerging studies on novel genome editing tools are focused on transgene-free editing, which are deemed to be more ‘regulatory-friendly’ and may attract improved public approval, and (4) the research publications are mostly focused on technical advancement in CRISPR systems (e.g., types of editing, targeting scope expansion, types of genomes targeted, and its delivery system) as portrayed by the frequency and relatedness of the extracted keywords in the network map (Figure 3).
Keywords related to regulatory, biosecurity, policymaking, and public acceptance issues are not present in the 50 most used keywords. Keywords related to climate change were also absent from the extracted sections. Despite this, climate change is one of the main driving forces for agricultural innovations to improve food sustainability and security. Regulatory approval and public opinion are also among the key deciding factors for genome-edited plants or crop adaption and commercialization [165,188]. The expanding gap between the fast-paced advances of CRISPR/Cas systems and the surrounding issues related to its regulation, adoption, and public acceptance should not be neglected if the potential of the technology in agriculture and food production is to be fully realized.
The present status of CRISPR technology allows for a wide range of applications aimed at increasing plant yield, disease resistance, and resilience to environmental changes. However, various technological advancements are still required, including precise editing and direct delivery of gene engineering reagents. One way to improve CRISPR delivery into the host genome is to reduce the cargo capacity so that a smaller delivery vehicle can be used to transfer CRISPR proteins through a cell. Another possibility is to use a hypercompact CRISPR CasΦ system. The CasΦ protein (~70 kD) has a molecular weight of half that of Cas9 and Cas12a enzymes [189]. Similar to Cas12a, CasΦ does not require a tracrRNA and produces staggered 5′-overhangs. It also has a minimal PAM requirement, allowing a wider range of target sites in the genome. Despite its low editing efficiency (0.85%), the possibility of using a hypercompact Cas delivery system may pave the way for the use of efficient small gene editors, further expanding the CRISPR editing toolbox.
The bibliographic analysis indicates that the trend of CRISPR/Cas research for the past decade has focused on various ways to improve the CRISPR/Cas functionality (e.g., targeting scope and delivery system). However, only recently, ‘natural brakes’ that could switch off the CRISPR/Cas activity when needed have been discovered. These tools are known as ‘anti-CRISPR’ technology, which uses phages and other mobile genetic elements that express anti-CRISPR proteins (Acrs). These proteins may nullify CRISPR/Cas activity by blocking Cas from binding or cleaving nucleic acid substrates [190]. Future improvements on these ‘natural brakes’ allow for more customized control of plant genome editing and expression, a needed innovation to improve the robustness of the existing CRISPR/Cas toolbox.
The recent development of biotechnologies and the production of novel crop varieties may benefit agricultural efficiency in the face of climate change. Establishing a technology adoption system across multiple farmlands is important to fully realize the potential benefits of these technologies and crop varieties. One of the issues towards adopting technology is the insufficient baseline empirical data to model the risks and benefits of sustainable farming across multiple farm types, farm sizes, and environments [191]. As technological developments are rapidly evolving, there is a constant need to deliver broad knowledge of sustainable farming to the public or industry to reduce the uncertainty about biotechnology and facilitate the adoption of agricultural biotechnology. These combined efforts will hopefully bring a paradigm shift in the farmer’s perspective on sustainable farming and work towards the common goal of food security.

Author Contributions

Conceptualization, B.C.T. and M.F.H.; writing—original draft preparation, M.F.H., C.K.S.K., E.Y.T., S.-E.L. and B.C.T.; funding acquisition, B.C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundamental Research Grant Scheme (FRGS/1/2018/STG03/UM/02/2) from the Ministry of Higher Education Malaysia, Royal Society-Newton Advanced Fellowship (NA170200; IF004-2018) from the Academy of Sciences Malaysia, and RU Fund (ST003-2021; RU004A-2020) from the Universiti Malaya.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ozdemir, D. The impact of climate change on agricultural productivity in Asian countries: A heterogeneous panel data approach. Environ. Sci. Pollut. Res. 2022, 29, 8205–8217. [Google Scholar] [CrossRef] [PubMed]
  2. Kuyah, S.; Sileshi, G.; Libère, N.; Chirinda, N.; Ndayisaba, P.; Dimobe, K.; Oborn, I. Innovative agronomic practices for sustainable intensification in sub-Saharan Africa. A review. Agron. Sustain. Dev. 2021, 41, 16. [Google Scholar] [CrossRef]
  3. Brás, T.; Seixas, J.; Carvalhais, N.; Jägermeyr, J. Severity of drought and heatwave crop losses tripled over the last five decades in Europe. Environ. Res. Lett. 2021, 16, 065012. [Google Scholar] [CrossRef]
  4. EC. Short-Term Outlook for EU Agricultural Markets in 2018 and 2019. Available online: https://agriculture.ec.europa.eu/data-and-analysis/markets/outlook/short-term_en (accessed on 1 May 2022).
  5. Yim, H. Incorporating climate change adaptation into sustainable development. J. Int. Dev. Coop. 2017, 2017, 139–171. [Google Scholar] [CrossRef]
  6. Galluzzi, G.; Seyoum, A.; Halewood, M.; Lopez Noriega, I.; Welch, E.W. The role of genetic resources in breeding for climate change: The case of public breeding programmes in eighteen developing countries. Plants 2020, 9, 1129. [Google Scholar] [CrossRef]
  7. Bandt, O.d.; Jacolin, L.; Lemaire, T. Climate Change in Developing Countries: Global Warming Effects, Transmission Channels and Adaptation Policies; Banque de France: Paris, France, 2021; pp. 1–65. [Google Scholar]
  8. Ludwig, F.; Terwisscha van Scheltinga, C.; Verhagen, J.; Kruijt, B.; van Ierland, E.; Dellink, R.; de Bruin, H.; Kabat, P. Climate Change Impacts on Developing Countries—EU Accountability; Think Tank Publications: Bulgaria, The Netherlands; pp. 1–45.
  9. FAO; IFAD; UNICEF; WFP; WHO. The State of Food Security and Nutrition in the World 2021: Transforming Food Systems for Food Security, Improved Nutrition and Affordable Healthy Diets for All; World Health Organization: Rome, Italy, 2021; pp. 1–240.
  10. UN. Policy Brief: The Impact of COVID-19 on Food Security and Nutrition; United Nations: Rome, Italy, 2020; pp. 1–22. [Google Scholar]
  11. Grassini, P.; Eskridge, K.M.; Cassman, K.G. Distinguishing between yield advances and yield plateaus in historical crop production trends. Nat. Commun. 2013, 4, 2918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Lau, S.-E.; Teo, W.F.A.; Teoh, E.Y.; Tan, B.C. Microbiome engineering and plant biostimulants for sustainable crop improvement and mitigation of biotic and abiotic stresses. Discov. Food 2022, 2, 9. [Google Scholar] [CrossRef]
  13. 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]
  14. Oshunsanya, S.O.; Nwosu, N.J.; Li, Y. Abiotic stress in agricultural crops under climatic conditions. In Sustainable Agriculture, Forest and Environmental Management; Jhariya, M.K., Banerjee, A., Meena, R.S., Yadav, D.K., Eds.; Springer Singapore: Singapore, 2019; pp. 71–100. [Google Scholar]
  15. Gao, C. Genome engineering for crop improvement and future agriculture. Cell 2021, 184, 1621–1635. [Google Scholar] [CrossRef]
  16. Lyzenga, W.J.; Pozniak, C.J.; Kagale, S. Advanced domestication: Harnessing the precision of gene editing in crop breeding. Plant Biotech. J. 2021, 19, 660–670. [Google Scholar] [CrossRef]
  17. Chaudhary, J.; Deshmukh, R.; Sonah, H. Mutagenesis approaches and their role in crop improvement. Plants 2019, 8, 467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Zhang, J. Important genomic regions mutate less often than do other regions. Nature 2022, 602, 38–39. [Google Scholar] [CrossRef] [PubMed]
  19. Ma, L.; Kong, F.; Sun, K.; Wang, T.; Guo, T. From classical radiation to modern radiation: Past, present, and future of radiation mutation breeding. Front. Public Health 2021, 9, 11. [Google Scholar] [CrossRef]
  20. Sedeek, K.E.M.; Mahas, A.; Mahfouz, M. Plant genome engineering for targeted improvement of crop traits. Front. Plant Sci. 2019, 10, 16. [Google Scholar] [CrossRef] [Green Version]
  21. Herman, R.A.; Fedorova, M.; Storer, N.P. Will following the regulatory script for GMOs promote public acceptance of gene-edited crops? Trends Biotechnol. 2019, 37, 1272–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Turnbull, C.; Lillemo, M.; Hvoslef-Eide, T.A.K. Global regulation of genetically modified crops amid the gene edited crop boom—A review. Front. Plant Sci. 2021, 12, 19. [Google Scholar] [CrossRef]
  23. 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]
  24. Chen, K.; Wang, Y.; Zhang, R.; Zhang, H.; Gao, C. CRISPR/Cas genome editing and precision plant breeding in agriculture. Annu. Rev. Plant Biol. 2019, 70, 667–697. [Google Scholar] [CrossRef]
  25. Wada, N.; Ueta, R.; Osakabe, Y.; Osakabe, K. Precision genome editing in plants: State-of-the-art in CRISPR/Cas9-based genome engineering. BMC Plant Biol. 2020, 20, 234. [Google Scholar] [CrossRef]
  26. Kashtwari, M.; Mansoor, S.; Wani, A.A.; Najar, M.A.; Deshmukh, R.K.; Baloch, F.S.; Abidi, I.; Zargar, S.M. Random mutagenesis in vegetatively propagated crops: Opportunities, challenges and genome editing prospects. Mol. Biol. Rep. 2021, 49, 5729–5749. [Google Scholar] [CrossRef]
  27. Silva, G.; Poirot, L.; Galetto, R.; Smith, J.; Montoya, G.; Duchateau, P.; Paques, F. Meganucleases and other tools for targeted genome engineering: Perspectives and challenges for gene therapy. Curr. Gene Ther. 2011, 11, 11–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Faye, G.; Dennebouy, N.; Kujawa, C.; Jacq, C. Inserted sequence in the mitochondrial 23s ribosomal RNA gene of the yeast Saccharomyces cerevisiae. Mol. Genet. Genomics 1979, 168, 101–109. [Google Scholar] [CrossRef] [PubMed]
  29. Bhambhani, S.; Kondhare, K.; Giri, A. Advanced genome editing strategies for manipulation of plant specialized metabolites pertaining to biofortification. Phytochem. Rev. 2022, 21, 81–99. [Google Scholar] [CrossRef]
  30. Iqbal, Z.; Iqbal, M.S.; Ahmad, A.; Memon, A.G.; Ansari, M.I. New prospects on the horizon: Genome editing to engineer plants for desirable traits. Curr. Plant Biol. 2020, 24, 100171. [Google Scholar] [CrossRef]
  31. Puchta, H.; Dujon, B.; Hohn, B. Homologous recombination in plant cells is enhanced by in vivo induction of double strand breaks into DNA by a site-specific endonuclease. Nucleic Acids Res. 1993, 21, 5034–5040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. D’Halluin, K.; Vanderstraeten, C.; Van Hulle, J.; Rosolowska, J.; Van Den Brande, I.; Pennewaert, A.; D’Hont, K.; Bossut, M.; Jantz, D.; Ruiter, R.; et al. Targeted molecular trait stacking in cotton through targeted double-strand break induction. Plant Biotechnol. J. 2013, 11, 933–941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Bijlani, S.; Pang, K.M.; Sivanandam, V.; Singh, A.; Chatterjee, S. The role of recombinant AAV in precise genome editing. Front. Genom. Editing 2021, 3, 799722. [Google Scholar] [CrossRef]
  34. Beyer, H.M.; Iwai, H. Structural basis for the propagation of homing endonuclease-associated inteins. Front. Mol. Biosci. 2022, 9, 14. [Google Scholar] [CrossRef]
  35. Shen, H.; Li, Z. DNA double-strand break repairs and their application in plant DNA integration. Genes 2022, 13, 322. [Google Scholar] [CrossRef]
  36. Lee, J.G.; Sung, Y.H.; Baek, I.J. Generation of genetically-engineered animals using engineered endonucleases. Arch. Pharm. Res. 2018, 41, 885–897. [Google Scholar] [CrossRef]
  37. Bonawitz, N.D.; Ainley, W.M.; Itaya, A.; Chennareddy, S.R.; Cicak, T.; Effinger, K.; Jiang, K.; Mall, T.K.; Marri, P.R.; Samuel, J.P.; et al. Zinc finger nuclease-mediated targeting of multiple transgenes to an endogenous soybean genomic locus via non-homologous end joining. Plant Biotechnol. J. 2019, 17, 750–761. [Google Scholar] [CrossRef] [PubMed]
  38. Shukla, V.K.; Doyon, Y.; Miller, J.C.; DeKelver, R.C.; Moehle, E.A.; Worden, S.E.; Mitchell, J.C.; Arnold, N.L.; Gopalan, S.; Meng, X.; et al. Precise genome modification in the crop species Zea mays using zinc-finger nucleases. Nature 2009, 459, 437–441. [Google Scholar] [CrossRef] [PubMed]
  39. Ran, Y.; Patron, N.; Kay, P.; Wong, D.; Buchanan, M.; Cao, Y.Y.; Sawbridge, T.; Davies, J.P.; Mason, J.; Webb, S.R.; et al. Zinc finger nuclease-mediated precision genome editing of an endogenous gene in hexaploid bread wheat (Triticum aestivum) using a DNA repair template. Plant Biotechnol. J. 2018, 16, 2088–2101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Jung, Y.-J.; Nogoy, F.M.; Lee, S.-K.; Cho, Y.-G.; Kang, K. Application of ZFN for site directed mutagenesis of rice SSIVa gene. Biotechnol. Bioproc. E 2018, 23, 108–115. [Google Scholar] [CrossRef]
  41. Rasheed, A.; Gill, R.A.; Hassan, M.U.; Mahmood, A.; Qari, S.; Zaman, Q.U.; Ilyas, M.; Aamer, M.; Batool, M.; Li, H.; et al. A critical review: Recent advancements in the use of CRISPR/Cas9 technology to enhance crops and alleviate global food crises. Curr. Issues Mol. Bio. 2021, 43, 1950–1976. [Google Scholar] [CrossRef]
  42. Romer, P.; Hahn, S.; Jordan, T.; Strauss, T.; Bonas, U.; Lahaye, T. Plant pathogen recognition mediated by promoter activation of the pepper Bs3 resistance gene. Science 2007, 318, 645–648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kay, S.; Hahn, S.; Marois, E.; Hause, G.; Bonas, U. A bacterial effector acts as a plant transcription factor and induces a cell size regulator. Science 2007, 318, 648–651. [Google Scholar] [CrossRef] [Green Version]
  44. Boch, J.; Scholze, H.; Schornack, S.; Landgraf, A.; Hahn, S.; Kay, S.; Lahaye, T.; Nickstadt, A.; Bonas, U. Breaking the code of DNA binding specificity of TAL-type III effectors. Science 2009, 326, 1509–1512. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Massel, K.; Godwin, I.D.; Gao, C. Applications and potential of genome editing in crop improvement. Genom. Biol. 2018, 19, 210. [Google Scholar] [CrossRef] [Green Version]
  46. Martinez-Fortun, J.; Phillips, D.W.; Jones, H.D. Potential impact of genome editing in world agriculture. Emerg. Top. Life Sci. 2017, 1, 117–133. [Google Scholar] [CrossRef]
  47. Li, T.; Liu, B.; Spalding, M.H.; Weeks, D.P.; Yang, B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nat. Biotechnol. 2012, 30, 390–392. [Google Scholar] [CrossRef] [PubMed]
  48. Shan, Q.; Zhang, Y.; Chen, K.; Zhang, K.; Gao, C. Creation of fragrant rice by targeted knockout of the OsBADH2 gene using TALEN technology. Plant Biotechnol. J. 2015, 13, 791–800. [Google Scholar] [CrossRef] [PubMed]
  49. 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]
  50. Demorest, Z.L.; Coffman, A.; Baltes, N.J.; Stoddard, T.J.; Clasen, B.M.; Luo, S.; Retterath, A.; Yabandith, A.; Gamo, M.E.; Bissen, J.; et al. Direct stacking of sequence-specific nuclease-induced mutations to produce high oleic and low linolenic soybean oil. BMC Plant Biol. 2016, 16, 225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Čermák, T.; Baltes, N.J.; Čegan, R.; Zhang, Y.; Voytas, D.F. High-frequency, precise modification of the tomato genome. Genome Biol. 2015, 16, 232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ishino, Y.; Shinagawa, H.; Makino, K.; Amemura, M.; Nakata, A. Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. J. Bacteriol. 1987, 169, 5429–5433. [Google Scholar] [CrossRef] [Green Version]
  53. Mojica, F.J.; Ferrer, C.; Juez, G.; Rodriguez-Valera, F. Long stretches of short tandem repeats are present in the largest replicons of the Archaea haloferax mediterranei and Haloferax volcanii and could be involved in replicon partitioning. Mol. Microbiol. 1995, 17, 85–93. [Google Scholar] [CrossRef]
  54. Jansen, R.; Embden, J.D.; Gaastra, W.; Schouls, L.M. Identification of genes that are associated with DNA repeats in prokaryotes. Mol. Microbiol. 2002, 43, 1565–1575. [Google Scholar] [CrossRef]
  55. Mojica, F.J.; Diez-Villasenor, C.; Garcia-Martinez, J.; Soria, E. Intervening sequences of regularly spaced prokaryotic repeats derive from foreign genetic elements. J. Mol. Evol. 2005, 60, 174–182. [Google Scholar] [CrossRef]
  56. Pourcel, C.; Salvignol, G.; Vergnaud, G. CRISPR elements in Yersinia pestis acquire new repeats by preferential uptake of bacteriophage DNA, and provide additional tools for evolutionary studies. Microbiology (Reading) 2005, 151, 653–663. [Google Scholar] [CrossRef]
  57. Sorek, R.; Lawrence, C.M.; Wiedenheft, B. CRISPR-mediated adaptive immune systems in bacteria and archaea. Annu. Rev. Biochem. 2013, 82, 237–266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Makarova, K.S.; Grishin, N.V.; Shabalina, S.A.; Wolf, Y.I.; Koonin, E.V. A putative RNA-interference-based immune system in prokaryotes: Computational analysis of the predicted enzymatic machinery, functional analogies with eukaryotic RNAi, and hypothetical mechanisms of action. Biol. Direct 2006, 1, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Barrangou, R.; Fremaux, C.; Deveau, H.; Richards, M.; Boyaval, P.; Moineau, S.; Romero, D.A.; Horvath, P. CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315, 1709–1712. [Google Scholar] [CrossRef] [PubMed]
  60. Deveau, H.; Barrangou, R.; Garneau, J.E.; Labonte, J.; Fremaux, C.; Boyaval, P.; Romero, D.A.; Horvath, P.; Moineau, S. Phage response to CRISPR-encoded resistance in Streptococcus thermophilus. J. Bacteriol. 2008, 190, 1390–1400. [Google Scholar] [CrossRef] [Green Version]
  61. Horvath, P.; Romero, D.A.; Coute-Monvoisin, A.C.; Richards, M.; Deveau, H.; Moineau, S.; Boyaval, P.; Fremaux, C.; Barrangou, R. Diversity, activity, and evolution of CRISPR loci in Streptococcus thermophilus. J. Bacteriol. 2008, 190, 1401–1412. [Google Scholar] [CrossRef] [Green Version]
  62. Karlson, C.K.S.; Mohd-Noor, S.N.; Nolte, N.; Tan, B.C. CRISPR/dCas9-based systems: Mechanisms and applications in plant sciences. Plants 2021, 10, 2055. [Google Scholar] [CrossRef]
  63. Makarova, K.S.; Wolf, Y.I.; Iranzo, J.; Shmakov, S.A.; Alkhnbashi, O.S.; Brouns, S.J.J.; Charpentier, E.; Cheng, D.; Haft, D.H.; Horvath, P.; et al. Evolutionary classification of CRISPR-Cas systems: A burst of class 2 and derived variants. Nat. Rev. Microbiol. 2020, 18, 67–83. [Google Scholar] [CrossRef]
  64. Zhang, Y.; Malzahn, A.A.; Sretenovic, S.; Qi, Y. The emerging and uncultivated potential of CRISPR technology in plant science. Nat. Plant. 2019, 5, 778–794. [Google Scholar] [CrossRef]
  65. Chaudhuri, A.; Halder, K.; Datta, A. Classification of CRISPR/Cas system and its application in tomato breeding. Theor. Appl. Genet. 2022, 135, 367–387. [Google Scholar] [CrossRef]
  66. Talas, A.; Simon, D.A.; Kulcsar, P.I.; Varga, E.; Krausz, S.L.; Welker, E. Bear reveals that increased fidelity variants can successfully reduce the mismatch tolerance of adenine but not cytosine base editors. Nat. Commun. 2021, 12, 6353. [Google Scholar] [CrossRef]
  67. Wu, S.; Li, L.; Li, M.; Sun, S.; Zhao, Y.; Xue, X.; Chen, F.; Zhong, J.; Guo, J.; Qu, Q.; et al. Two compact Cas9 ortholog-based cytosine base editors expand the DNA targeting scope and applications in vitro and in vivo. Front. Cell. Dev. Biol. 2022, 10, 809922. [Google Scholar] [CrossRef]
  68. Wang, G.; Xu, Z.; Wang, F.; Huang, Y.; Xin, Y.; Liang, S.; Li, B.; Si, H.; Sun, L.; Wang, Q.; et al. Development of an efficient and precise adenine base editor (ABE) with expanded target range in allotetraploid cotton (Gossypium hirsutum). BMC Biol. 2022, 20, 45. [Google Scholar] [CrossRef]
  69. 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] [PubMed] [Green Version]
  70. 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] [PubMed] [Green Version]
  71. Molla, K.A.; Yang, Y. CRISPR/Cas-mediated base editing: Technical considerations and practical applications. Trends Biotechnol. 2019, 37, 1121–1142. [Google Scholar] [CrossRef] [PubMed]
  72. Li, Q.; Li, Y.; Yang, S.; Huang, S.; Yan, M.; Ding, Y.; Tang, W.; Lou, X.; Yin, Q.; Sun, Z.; et al. CRISPR-Cas9-mediated base-editing screening in mice identifies DND1 amino acids that are critical for primordial germ cell development. Nat. Cell Biol. 2018, 20, 1315–1325. [Google Scholar] [CrossRef]
  73. 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]
  74. Kang, B.C.; Yun, J.Y.; Kim, S.T.; Shin, Y.; Ryu, J.; Choi, M.; Woo, J.W.; Kim, J.S. Precision genome engineering through adenine base editing in plants. Nat. Plants 2018, 4, 427–431. [Google Scholar] [CrossRef]
  75. Scholefield, J.; Harrison, P.T. Prime editing—An update on the field. Gene Ther. 2021, 28, 396–401. [Google Scholar] [CrossRef]
  76. Jiang, Y.Y.; Chai, Y.P.; Lu, M.H.; Han, X.L.; Lin, Q.; Zhang, Y.; Zhang, Q.; Zhou, Y.; Wang, X.C.; Gao, C.; et al. Prime editing efficiently generates W542L and S621I double mutations in two ALS genes in maize. Genome Biol. 2020, 21, 257. [Google Scholar] [CrossRef]
  77. Li, H.; Li, J.; Chen, J.; Yan, L.; Xia, L. Precise modifications of both exogenous and endogenous genes in rice by prime editing. Mol. Plant 2020, 13, 671–674. [Google Scholar] [CrossRef] [PubMed]
  78. Perotti, V.E.; Larran, A.S.; Palmieri, V.E.; Martinatto, A.K.; Alvarez, C.E.; Tuesca, D.; Permingeat, H.R. A novel triple amino acid substitution in the EPSPS found in a high-level glyphosate-resistant Amaranthus hybridus population from Argentina. Pest Manag. Sci. 2019, 75, 1242–1251. [Google Scholar] [CrossRef] [PubMed]
  79. Ha, D.I.; Lee, J.M.; Lee, N.E.; Kim, D.; Ko, J.H.; Kim, Y.S. Highly efficient and safe genome editing by CRISPR-Cas12a using CRISPR RNA with a ribosyl-2’-o-methylated uridinylate-rich 3’-overhang in mouse zygotes. Exp. Mol. Med. 2020, 52, 1823–1830. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, Y.; Wang, M.; Zheng, T.; Hou, Y.; Zhang, P.; Tang, T.; Wei, J.; Du, Q. Specificity profiling of CRISPR system reveals greatly enhanced off-target gene editing. Sci. Rep. 2020, 10, 2269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Swarts, D.C.; van der Oost, J.; Jinek, M. Structural basis for guide RNA processing and seed-dependent DNA targeting by CRISPR-Cas12a. Mol. Cell 2017, 66, 221–233.e4. [Google Scholar] [CrossRef] [Green Version]
  82. Mahas, A.; Marsic, T.; Lopez-Portillo Masson, M.; Wang, Q.; Aman, R.; Zheng, C.; Ali, Z.; Alsanea, M.; Al-Qahtani, A.; Ghanem, B.; et al. Characterization of a thermostable CAS13 enzyme for one-pot detection of SARS-COV-2. Proc. Natl. Acad. Sci. USA 2022, 119, e2118260119. [Google Scholar] [CrossRef]
  83. Bayoumi, M.; Munir, M. Potential use of CRISPR/cas13 machinery in understanding virus–host interaction. Front. Microbiol. 2021, 12, 3580. [Google Scholar] [CrossRef]
  84. Savage, D.F. Cas14: Big advances from small CRISPR proteins. Biochemistry 2019, 58, 1024–1025. [Google Scholar] [CrossRef] [Green Version]
  85. Badhan, S.; Ball, A.S.; Mantri, N. First report of CRISPR/Cas9 mediated DNA-free editing of 4CL and RVE7 genes in chickpea protoplasts. Int. J. Mol. Sci. 2021, 22, 396. [Google Scholar] [CrossRef]
  86. Zeng, Y.; Wen, J.; Zhao, W.; Wang, Q.; Huang, W. Rational improvement of rice yield and cold tolerance by editing the three genes OsPIN5b, GS3, and OsMYB30 with the CRISPR-Cas9 system. Front. Plant Sci. 2019, 10, 1663. [Google Scholar] [CrossRef]
  87. Li, Y.; Zhu, J.; Wu, H.; Liu, C.; Huang, C.; Lan, J.; Zhao, Y.; Xie, C. Precise base editing of non-allelic acetolactate synthase genes confers sulfonylurea herbicide resistance in maize. Crop J. 2020, 8, 449–456. [Google Scholar] [CrossRef]
  88. Bouzroud, S.; Gasparini, K.; Hu, G.; Antonia, M.; Barbosa, M.; Luan Rosa, B.; Fahr, M.; Bendaou, N.; Bouzayen, M.; Zsögön, A.; et al. Down regulation and loss of auxin response factor 4 function using CRISPR/Cas9 alters plant growth, stomatal function and improves tomato tolerance to salinity and osmotic stress. Genes 2020, 11, 272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chen, Q.; Tang, W.; Zeng, G.; Sheng, H.; Shi, W.; Xiao, Y. Reduction of cadmium accumulation in the grains of male sterile rice Chuang-5S carrying Pi48 or Pi49 through marker-assisted selection. 3 Biotech 2020, 10, 539. [Google Scholar] [CrossRef]
  90. Yu, W.; Wang, L.; Zhao, R.; Sheng, J.; Zhang, S.; Li, R.; Shen, L. Knockout of SlMAPK3 enhances tolerance to heat stress involving ROS homeostasis in tomato plants. BMC Plant Biol. 2019, 19, 354. [Google Scholar] [CrossRef] [Green Version]
  91. Kis, A.; Hamar, E.; Tholt, G.; Ban, R.; Havelda, Z. Creating highly efficient resistance against wheat dwarf virus in barley by employing CRISPR/Cas9 system. Plant Biotechnol. J. 2019, 17, 1004–1006. [Google Scholar] [CrossRef]
  92. Foster, A.J.; Martin-Urdiroz, M.; Yan, X.; Wright, H.S.; Soanes, D.M.; Talbot, N.J. CRISPR-Cas9 ribonucleoprotein-mediated co-editing and counterselection in the rice blast fungus. Sci. Rep. 2018, 8, 14355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. 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] [PubMed] [Green Version]
  94. Zhang, Y.; Guo, W.; Chen, L.; Shen, X.; Yang, H.; Fang, Y.; Ouyang, W.; Mai, S.; Chen, H.; Chen, S.; et al. CRISPR/Cas9-mediated targeted mutagenesis of GmUGT enhanced soybean resistance against leaf-chewing insects through flavonoids biosynthesis. Front. Plant Sci. 2022, 13, 802716. [Google Scholar] [CrossRef]
  95. Miao, C.; Xiao, L.; Hua, K.; Zou, C.; Zhao, Y.; Bressan, R.A.; Zhu, J.-K. Mutations in a subfamily of abscisic acid receptor genes promote rice growth and productivity. Proc. Natl. Acad. Sci. USA 2018, 115, 6058–6063. [Google Scholar] [CrossRef] [Green Version]
  96. Zhang, Z.; Hua, L.; Gupta, A.; Tricoli, D.; Edwards, K.J.; Yang, B.; Li, W. Development of an Agrobacterium-delivered CRISPR/Cas9 system for wheat genome editing. Plant Biotechnol. J. 2019, 17, 1623–1635. [Google Scholar] [CrossRef]
  97. Dong, O.X.; Yu, S.; Jain, R.; Zhang, N.; Duong, P.Q.; Butler, C.; Li, Y.; Lipzen, A.; Martin, J.A.; Barry, K.W.; et al. Marker-free carotenoid-enriched rice generated through targeted gene insertion using CRISPR-Cas9. Nat. Commun. 2020, 11, 1178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Zhou, J.; Xin, X.; He, Y.; Chen, H.; Li, Q.; Tang, X.; Zhong, Z.; Deng, K.; Zheng, X.; Akher, S.A.; et al. Multiplex QTL editing of grain-related genes improves yield in elite rice varieties. Plant Cell Rep. 2019, 38, 475–485. [Google Scholar] [CrossRef] [PubMed]
  99. Yuyu, C.; Aike, Z.; Pao, X.; Xiaoxia, W.; Yongrun, C.; Beifang, W.; Yue, Z.; Liaqat, S.; Shihua, C.; Liyong, C.; et al. Effects of GS3 and GL3.1 for grain size editing by CRISPR/Cas9 in rice. Rice Sci. 2020, 27, 405–413. [Google Scholar] [CrossRef]
  100. Yuste-Lisbona, F.J.; Fernandez-Lozano, A.; Pineda, B.; Bretones, S.; Ortiz-Atienza, A.; Garcia-Sogo, B.; Muller, N.A.; Angosto, T.; Capel, J.; Moreno, V.; et al. ENO regulates tomato fruit size through the floral meristem development network. Proc. Natl. Acad. Sci. USA 2020, 117, 8187–8195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Neupane, D.; Adhikari, P.; Bhattarai, D.; Rana, B.; Ahmed, Z.; Sharma, U.; Adhikari, D. Does climate change affect the yield of the top three cereals and food security in the world? Earth 2022, 3, 45–71. [Google Scholar] [CrossRef]
  102. Lau, S.E.; Hamdan, M.F.; Pua, T.L.; Saidi, N.B.; Tan, B.C. Plant nitric oxide signaling under drought stress. Plants 2021, 10, 360. [Google Scholar] [CrossRef]
  103. Mohd Amnan, M.A.; Pua, T.L.; Lau, S.E.; Tan, B.C.; Yamaguchi, H.; Hitachi, K.; Tsuchida, K.; Komatsu, S. Osmotic stress in banana is relieved by exogenous nitric oxide. PeerJ 2021, 9, e10879. [Google Scholar] [CrossRef]
  104. Vats, S.; Kumawat, S.; Kumar, V.; Patil, G.B.; Joshi, T.; Sonah, H.; Sharma, T.R.; Deshmukh, R. Genome editing in plants: Exploration of technological advancements and challenges. Cell 2019, 8, 1386. [Google Scholar] [CrossRef] [Green Version]
  105. Zhang, A.; Liu, Y.; Wang, F.; Li, T.; Chen, Z.; Kong, D.; Bi, J.; Zhang, F.; Luo, X.; Wang, J. Enhanced rice salinity tolerance via CRISPR/Cas9-targeted mutagenesis of the OsRR22 gene. Mol. Breed. 2019, 39, 47. [Google Scholar] [CrossRef] [Green Version]
  106. Bo, W.; Zhaohui, Z.; Huanhuan, Z.; Xia, W.; Binglin, L.; Lijia, Y.; Xiangyan, H.; Deshui, Y.; Xuelian, Z.; Chunguo, W.; et al. Targeted mutagenesis of NAC transcription factor gene, OsNAC041, leading to salt sensitivity in rice. Rice Sci. 2019, 26, 98–108. [Google Scholar] [CrossRef]
  107. Xie, Z.; Nolan, T.M.; Jiang, H.; Yin, Y. AP2/ERF transcription factor regulatory networks in hormone and abiotic stress responses in Arabidopsis. Front. Plant Sci. 2019, 10, 228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Faraji, S.; Filiz, E.; Kazemitabar, S.K.; Vannozzi, A.; Palumbo, F.; Barcaccia, G.; Heidari, P. The AP2/ERF gene family in Triticum durum: Genome-wide identification and expression analysis under drought and salinity stresses. Genes 2020, 11, 1464. [Google Scholar] [CrossRef] [PubMed]
  109. Duan, Y.B.; Li, J.; Qin, R.Y.; Xu, R.F.; Li, H.; Yang, Y.C.; Ma, H.; Li, L.; Wei, P.C.; Yang, J.B. Identification of a regulatory element responsible for salt induction of rice OsRAV2 through ex situ and in situ promoter analysis. Plant Mol. Bio. 2016, 90, 49–62. [Google Scholar] [CrossRef] [PubMed]
  110. Liu, X.; Wu, D.; Shan, T.; Xu, S.; Qin, R.; Li, H.; Negm, M.; Wu, D.; Li, J. The trihelix transcription factor OsGT γ-2 is involved adaption to salt stress in rice. Plant Mol. Bio. 2020, 103, 545–560. [Google Scholar] [CrossRef]
  111. Nazir, R.; Mandal, S.; Mitra, S.; Ghorai, M.; Das, N.; Jha, N.K.; Majumder, M.; Pandey, D.K.; Dey, A. Clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated genome-editing toolkit to enhance salt stress tolerance in rice and wheat. Physiol. Plant. 2022, 174, 12. [Google Scholar] [CrossRef]
  112. Wang, T.; Xun, H.; Wang, W.; Ding, X.; Tian, H.; Hussain, S.; Dong, Q.; Li, Y.; Cheng, Y.; Wang, C.; et al. Mutation of GmAITR genes by CRISPR/Cas9 genome editing results in enhanced salinity stress tolerance in soybean. Front. Plant Sci. 2021, 12, 779598. [Google Scholar] [CrossRef]
  113. Zhang, M.; Cao, Y.; Wang, Z.; Wang, Z.Q.; Shi, J.; Liang, X.; Song, W.; Chen, Q.; Lai, J.; Jiang, C. A retrotransposon in an HKT1 family sodium transporter causes variation of leaf Na(+) exclusion and salt tolerance in maize. New Phytol. 2018, 217, 1161–1176. [Google Scholar] [CrossRef] [Green Version]
  114. Tran, M.T.; Doan, D.T.H.; Kim, J.; Song, Y.J.; Sung, Y.W.; Das, S.; Kim, E.J.; Son, G.H.; Kim, S.H.; Van Vu, T.; et al. CRISPR/Cas9-based precise excision of SlHyPRP1 domain(s) to obtain salt stress-tolerant tomato. Plant Cell Rep. 2021, 40, 999–1011. [Google Scholar] [CrossRef]
  115. Amnan, M.A.M.; Aizat, W.M.; Khaidizar, F.D.; Tan, B.C. Drought stress induces morpho-physiological and proteome changes of Pandanus amaryllifolius. Plants 2022, 11, 221. [Google Scholar] [CrossRef]
  116. Zhang, Y.; Wang, X.; Luo, Y.; Zhang, L.; Yao, Y.; Han, L.; Chen, Z.; Wang, L.; Li, Y. OsABA8ox2, an ABA catabolic gene, suppresses root elongation of rice seedlings and contributes to drought response. Crop J. 2020, 8, 480–491. [Google Scholar] [CrossRef]
  117. Ogata, T.; Ishizaki, T.; Fujita, M.; Fujita, Y. CRISPR/Cas9-targeted mutagenesis of osera1 confers enhanced responses to abscisic acid and drought stress and increased primary root growth under nonstressed conditions in rice. PLoS ONE 2020, 15, e0243376. [Google Scholar] [CrossRef] [PubMed]
  118. Yin, X.; Biswal, A.K.; Dionora, J.; Perdigon, K.M.; Balahadia, C.P.; Mazumdar, S.; Chater, C.; Lin, H.C.; Coe, R.A.; Kretzschmar, T.; et al. CRISPR-Cas9 and CRISPR-Cpf1 mediated targeting of a stomatal developmental gene EPFL9 in rice. Plant Cell Rep. 2017, 36, 745–757. [Google Scholar] [CrossRef] [PubMed]
  119. Caine, R.S.; Yin, X.; Sloan, J.; Harrison, E.L.; Mohammed, U.; Fulton, T.; Biswal, A.K.; Dionora, J.; Chater, C.C.; Coe, R.A.; et al. Rice with reduced stomatal density conserves water and has improved drought tolerance under future climate conditions. New Phytol. 2019, 221, 371–384. [Google Scholar] [CrossRef] [PubMed]
  120. Hughes, J.; Hepworth, C.; Dutton, C.; Dunn, J.A.; Hunt, L.; Stephens, J.; Waugh, R.; Cameron, D.D.; Gray, J.E. Reducing stomatal density in barley improves drought tolerance without impacting on yield. Plant Physiol. 2017, 174, 776–787. [Google Scholar] [CrossRef] [Green Version]
  121. Clemens, M.; Faralli, M.; Lagreze, J.; Bontempo, L.; Piazza, S.; Varotto, C.; Malnoy, M.; Oechel, W.; Rizzoli, A.; Dalla Costa, L. VvEPFL9-1 knock-out via CRISPR/Cas9 reduces stomatal density in grapevine. Front. Plant Sci. 2022, 13, 878001. [Google Scholar] [CrossRef]
  122. Wang, L.; Chen, L.; Li, R.; Zhao, R.; Yang, M.; Sheng, J.; Shen, L. Reduced drought tolerance by CRISPR/Cas9-mediated SlMAPK3 mutagenesis in tomato plants. J. Agric. Food Chem. 2017, 65, 8674–8682. [Google Scholar] [CrossRef]
  123. Nieves-Cordones, M.; Mohamed, S.; Tanoi, K.; Kobayashi, N.I.; Takagi, K.; Vernet, A.; Guiderdoni, E.; Perin, C.; Sentenac, H.; Very, A.A. Production of low-Cs+ rice plants by inactivation of the k+ transporter OsHAK1 with the CRISPR-Cas system. Plant J. 2017, 92, 43–56. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, F.Z.; Chen, M.X.; Yu, L.J.; Xie, L.J.; Yuan, L.B.; Qi, H.; Xiao, M.; Guo, W.; Chen, Z.; Yi, K.; et al. OsARM1, an R2R3 MYB transcription factor, is involved in regulation of the response to arsenic stress in rice. Front. Plant Sci. 2017, 8, 1868. [Google Scholar] [CrossRef] [Green Version]
  125. Tang, L.; Mao, B.; Li, Y.; Lv, Q.; Zhang, L.; Chen, C.; He, H.; Wang, W.; Zeng, X.; Shao, Y.; et al. Knockout of OsNramp5 using the CRISPR/Cas9 system produces low Cd-accumulating indica rice without compromising yield. Sci Rep. 2017, 7, 14438. [Google Scholar] [CrossRef] [Green Version]
  126. Shao, X.; Wu, S.; Dou, T.; Zhu, H.; Hu, C.; Huo, H.; He, W.; Deng, G.; Sheng, O.; Bi, F.; et al. Using CRISPR/Cas9 genome editing system to create MaGA20ox2 gene-modified semi-dwarf banana. Plant Biotechnol. J. 2020, 18, 17–19. [Google Scholar] [CrossRef]
  127. 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]
  128. Talakayala, A.; Ankanagari, S.; Garladinne, M. CRISPR-Cas genome editing system: A versatile tool for developing disease resistant crops. Plant Stress 2022, 3, 100056. [Google Scholar] [CrossRef]
  129. 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]
  130. Jiang, W.; Zhou, H.; Bi, H.; Fromm, M.; Yang, B.; Weeks, D.P. Demonstration of CRISPR/Cas9/sgRNA-mediated targeted gene modification in Arabidopsis, tobacco, sorghum and rice. Nucleic Acids Res. 2013, 41, e188. [Google Scholar] [CrossRef]
  131. 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]
  132. 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] [PubMed] [Green Version]
  133. 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] [PubMed] [Green Version]
  134. Macovei, A.; Sevilla, N.R.; Cantos, C.; Jonson, G.B.; Slamet-Loedin, I.; Čermák, T.; Voytas, D.F.; Choi, I.R.; Chadha-Mohanty, P. Novel alleles of rice eif4g generated by CRISPR/Cas9-targeted mutagenesis confer resistance to rice tungro spherical virus. Plant Biotechnol. J. 2018, 16, 1918–1927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. 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] [Green Version]
  136. Thomazella, D.P.d.T.; Seong, K.; Mackelprang, R.; Dahlbeck, D.; Geng, Y.; Gill, U.S.; Qi, T.; Pham, J.; Giuseppe, P.; Lee, C.Y.; et al. Loss of function of a dmr6 ortholog in tomato confers broad-spectrum disease resistance. Proc. Natl. Acad. Sci. USA 2021, 118, e2026152118. [Google Scholar] [CrossRef]
  137. Kumar, N.; Galli, M.; Ordon, J.; Stuttmann, J.; Kogel, K.-H.; Imani, J. Further analysis of barley MORC1 using a highly efficient RNA-guided Cas9 gene-editing system. Plant Biotechnol. J. 2018, 16, 1892–1903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Nekrasov, V.; Wang, C.; Win, J.; Lanz, C.; Weigel, D.; Kamoun, S. Rapid generation of a transgene-free powdery mildew resistant tomato by genome deletion. Sci. Rep. 2017, 7, 482. [Google Scholar] [CrossRef] [Green Version]
  139. 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] [PubMed] [Green Version]
  140. Wan, D.-Y.; Guo, Y.; Cheng, Y.; Hu, Y.; Xiao, S.; Wang, Y.; Wen, Y.-Q. CRISPR/Cas9-mediated mutagenesis of VvMLO3 results in enhanced resistance to powdery mildew in grapevine (Vitis vinifera). Hortic. Res. 2020, 7, 116. [Google Scholar] [CrossRef]
  141. Wang, Y.Y.; Zhang, Y.J.; Zhou, R.; Dossa, K.; Yu, J.Y.; Li, D.H.; Liu, A.L.; Mmadi, M.A.; Zhang, X.R.; You, J. Identification and characterization of the bzip transcription factor family and its expression in response to abiotic stresses in sesame. PLoS ONE 2018, 13, 21. [Google Scholar] [CrossRef] [PubMed]
  142. Wei, J.; Fang, Y.; Jiang, H.; Wu, X.T.; Zuo, J.H.; Xia, X.C.; Li, J.Q.; Stich, B.; Cao, H.; Liu, Y.X. Combining QTL mapping and gene co-expression network analysis for prediction of candidate genes and molecular network related to yield in wheat. BMC Plant Biol. 2022, 22, 288. [Google Scholar] [CrossRef] [PubMed]
  143. Song, G.; Jia, M.; Chen, K.; Kong, X.; Khattak, B.; Xie, C.; Li, A.; Mao, L. CRISPR/Cas9: A powerful tool for crop genome editing. Crop J. 2016, 4, 75–82. [Google Scholar] [CrossRef] [Green Version]
  144. Ma, X.; Zhu, Q.; Chen, Y.; Liu, Y.-G. CRISPR/Cas9 platforms for genome editing in plants: Developments and applications. Mol. Plant 2016, 9, 961–974. [Google Scholar] [CrossRef] [Green Version]
  145. 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]
  146. Xu, R.; Yang, Y.; Qin, R.; Li, H.; Qiu, C.; Li, L.; Wei, P.; Yang, J. Rapid improvement of grain weight via highly efficient CRISPR/Cas9-mediated multiplex genome editing in rice. J. Genet. Genomics 2016, 43, 529–532. [Google Scholar] [CrossRef]
  147. Zhang, Y.; Liang, Z.; Zong, Y.; Wang, Y.; Liu, J.; Chen, K.; Qiu, J.-L.; Gao, C. Efficient and transgene-free genome editing in wheat through transient expression of CRISPR/Cas9 DNA or RNA. Nat. Commun. 2016, 7, 12617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Rodriguez-Leal, D.; Lemmon, Z.H.; Man, J.; Bartlett, M.E.; Lippman, Z.B. Engineering quantitative trait variation for crop improvement by genome editing. Cell 2017, 171, 470–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Gao, H.; Gadlage, M.J.; Lafitte, H.R.; Lenderts, B.; Yang, M.; Schroder, M.; Farrell, J.; Snopek, K.; Peterson, D.; Feigenbutz, L.; et al. Superior field performance of waxy corn engineered using CRISPR-Cas9. Nat. Biotechnol. 2020, 38, 579–581. [Google Scholar] [CrossRef]
  150. Mao, H.; Jian, C.; Cheng, X.; Chen, B.; Mei, F.; Li, F.; Zhang, Y.; Li, S.; Du, L.; Li, T.; et al. The wheat ABA receptor gene TaPYL1-1b contributes to drought tolerance and grain yield by increasing water-use efficiency. Plant Biotechnol. J. 2022, 20, 846–861. [Google Scholar] [CrossRef] [PubMed]
  151. Soyk, S.; Muller, N.A.; Park, S.J.; Schmalenbach, I.; Jiang, K.; Hayama, R.; Zhang, L.; Van Eck, J.; Jimenez-Gomez, J.M.; Lippman, Z.B. Variation in the flowering gene self pruning 5g promotes day-neutrality and early yield in tomato. Nat. Genet. 2017, 49, 162–168. [Google Scholar] [CrossRef]
  152. Carmel-Goren, L.; Liu, Y.S.; Lifschitz, E.; Zamir, D. The self-pruning gene family in tomato. Plant Mol. Biol. 2003, 52, 1215–1222. [Google Scholar] [CrossRef]
  153. Klap, C.; Yeshayahou, E.; Bolger, A.M.; Arazi, T.; Gupta, S.K.; Shabtai, S.; Usadel, B.; Salts, Y.; Barg, R. Tomato facultative parthenocarpy results from slagamous-like 6 loss of function. Plant Biotechnol. J. 2017, 15, 634–647. [Google Scholar] [CrossRef]
  154. van Eck, N.J.; Waltman, L. Software survey: VOSviewer, a computer program for bibliometric mapping. Scientometrics 2010, 84, 523–538. [Google Scholar] [CrossRef] [Green Version]
  155. Romero, L.; Portillo-Salido, E. Trends in Sigma-1 receptor research: A 25-year bibliometric analysis. Front. Pharmacol. 2019, 10, 564. [Google Scholar] [CrossRef] [Green Version]
  156. Hamdan, M.F.; Mohd Noor, S.N.; Abd-Aziz, N.; Pua, T.-L.; Tan, B.C. Green revolution to gene revolution: Technological advances in agriculture to feed the world. Plants 2022, 11, 1297. [Google Scholar] [CrossRef]
  157. Liu, Z.; Dong, H.; Cui, Y.; Cong, L.; Zhang, D. Application of different types of CRISPR/Cas-based systems in bacteria. Microb. Cell Fact. 2020, 19, 172. [Google Scholar] [CrossRef] [PubMed]
  158. Zhang, Y.; Ren, Q.; Tang, X.; Liu, S.; Malzahn, A.A.; Zhou, J.; Wang, J.; Yin, D.; Pan, C.; Yuan, M.; et al. Expanding the scope of plant genome engineering with Cas12a orthologs and highly multiplexable editing systems. Nat. Commun. 2021, 12, 1944. [Google Scholar] [CrossRef] [PubMed]
  159. Huang, T.K.; Puchta, H. Novel CRISPR/Cas applications in plants: From prime editing to chromosome engineering. Transgenic Res. 2021, 30, 529–549. [Google Scholar] [CrossRef]
  160. Wada, N.; Osakabe, K.; Osakabe, Y. Expanding the plant genome editing toolbox with recently developed CRISPR-Cas systems. Plant Physiol. 2022, 188, 1825–1837. [Google Scholar] [CrossRef]
  161. Zess, E.; Begemann, M. CRISPR-Cas9 and beyond: What’s next in plant genome engineering. In Vitro Cell. Dev. Biol. Plant 2021, 57, 584–594. [Google Scholar] [CrossRef]
  162. Budeguer, F.; Enrique, R.; Perera, M.F.; Racedo, J.; Castagnaro, A.P.; Noguera, A.S.; Welin, B. Genetic transformation of sugarcane, current status and future prospects. Front. Plant Sci. 2021, 12, 20. [Google Scholar] [CrossRef]
  163. Lassoued, R.; Macall, D.M.; Hesseln, H.; Phillips, P.W.B.; Smyth, S.J. Benefits of genome-edited crops: Expert opinion. Transgenic Res. 2019, 28, 247–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Ahmad, S.; Tang, L.; Shahzad, R.; Mawia, A.M.; Rao, G.S.; Jamil, S.; Wei, C.; Sheng, Z.; Shao, G.; Wei, X.; et al. CRISPR-based crop improvements: A way forward to achieve zero hunger. J. Agric. Food Chem. 2021, 69, 8307–8323. [Google Scholar] [CrossRef]
  165. Entine, J.; Felipe, M.S.S.; Groenewald, J.H.; Kershen, D.L.; Lema, M.; McHughen, A.; Nepomuceno, A.L.; Ohsawa, R.; Ordonio, R.L.; Parrott, W.A.; et al. Regulatory approaches for genome edited agricultural plants in select countries and jurisdictions around the world. Transgenic Res. 2021, 30, 551–584. [Google Scholar] [CrossRef]
  166. Lobato-Gomez, M.; Hewitt, S.; Capell, T.; Christou, P.; Dhingra, A.; Giron-Calva, P.S. Transgenic and genome-edited fruits: Background, constraints, benefits, and commercial opportunities. Hortic. Res. 2021, 8, 166. [Google Scholar] [CrossRef]
  167. Anjanappa, R.B.; Gruissem, W. Current progress and challenges in crop genetic transformation. J. Plant Physiol. 2021, 261, 153411. [Google Scholar] [CrossRef] [PubMed]
  168. Zhao, H.; Jia, Y.; Cao, Y.; Wang, Y. Improving T-DNA transfer to Tamarix hispida by adding chemical compounds during Agrobacterium tumefaciens culture. Front. Plant Sci. 2020, 11, 501358. [Google Scholar] [CrossRef] [PubMed]
  169. Yue, J.J.; Yuan, J.L.; Wu, F.H.; Yuan, Y.H.; Cheng, Q.W.; Hsu, C.T.; Lin, C.S. Protoplasts: From isolation to CRISPR/Cas genome editing application. Front. Genom. Ed. 2021, 3, 717017. [Google Scholar] [CrossRef] [PubMed]
  170. Yoo, B.C.; Yadav, N.S.; Orozco, E.M.; Sakai, H. Cas9/gRNA-mediated genome editing of yeast mitochondria and Chlamydomonas chloroplasts. PeerJ 2020, 8, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Sandhya, D.; Jogam, P.; Allini, V.R.; Abbagani, S.; Alok, A. The present and potential future methods for delivering CRISPR/Cas9 components in plants. J. Genet. Eng. Biotechnol. 2020, 18, 25. [Google Scholar] [CrossRef] [PubMed]
  172. Hua, K.; Zhang, J.; Botella, J.; Ma, C.; Kong, F.; Liu, B.; Zhu, J. Perspectives on the application of genome-editing technologies in crop breeding. Mol. Plant 2019, 12, 1047–1059. [Google Scholar] [CrossRef] [Green Version]
  173. Chen, S.-J. Minimizing off-target effects in CRISPR-Cas9 genome editing. Cell Biol. Toxicol. 2019, 35, 399–401. [Google Scholar] [CrossRef] [Green Version]
  174. Kawall, K.; Cotter, J.; Then, C. Broadening the GMO risk assessment in the EU for genome editing technologies in agriculture. Environ. Sci. Eur. 2020, 32, 106. [Google Scholar] [CrossRef]
  175. Muller, M.; Lee, C.M.; Gasiunas, G.; Davis, T.H.; Cradick, T.J.; Siksnys, V.; Bao, G.; Cathomen, T.; Mussolino, C. Streptococcus thermophilus CRISPR-Cas9 systems enable specific editing of the human genome. Mol. Ther. 2016, 24, 636–644. [Google Scholar] [CrossRef] [Green Version]
  176. Nishimasu, H.; Shi, X.; Ishiguro, S.; Gao, L.; Hirano, S.; Okazaki, S.; Noda, T.; Abudayyeh, O.O.; Gootenberg, J.S.; Mori, H.; et al. Engineered CRISPR-Cas9 nuclease with expanded targeting space. Science 2018, 361, 1259–1262. [Google Scholar] [CrossRef]
  177. Chen, J.S.; Dagdas, Y.S.; Kleinstiver, B.P.; Welch, M.M.; Sousa, A.A.; Harrington, L.B.; Sternberg, S.H.; Joung, J.K.; Yildiz, A.; Doudna, J.A. Enhanced proofreading governs CRISPR–Cas9 targeting accuracy. Nature 2017, 550, 407–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Slaymaker, I.M.; Gao, L.Y.; Zetsche, B.; Scott, D.A.; Yan, W.X.; Zhang, F. Rationally engineered Cas9 nucleases with improved specificity. Science 2016, 351, 84–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Kleinstiver, B.P.; Pattanayak, V.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Zheng, Z.; Joung, J.K. High-fidelity CRISPR–Cas9 nucleases with no detectable genome-wide off-target effects. Nature 2016, 529, 490–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Dugar, G.; Leenay, R.T.; Eisenbart, S.K.; Bischler, T.; Aul, B.U.; Beisel, C.L.; Sharma, C.M. CRISPR RNA-dependent binding and cleavage of endogenous RNAs by the Campylobacter jejuni Cas9. Mol. Cell 2018, 69, 893–905. [Google Scholar] [CrossRef] [Green Version]
  181. Nakamura, M.; Gao, Y.; Dominguez, A.A.; Qi, L.S. CRISPR technologies for precise epigenome editing. Nat. Cell Biol. 2021, 23, 11–22. [Google Scholar] [CrossRef]
  182. Galonska, C.; Charlton, J.; Mattei, A.L.; Donaghey, J.; Clement, K.; Gu, H.; Mohammad, A.W.; Stamenova, E.K.; Cacchiarelli, D.; Klages, S.; et al. Genome-wide tracking of dCas9-methyltransferase footprints. Nat. Commun. 2018, 9, 597. [Google Scholar] [CrossRef] [Green Version]
  183. Lin, Q.; Zong, Y.; Xue, C.; Wang, S.; Jin, S.; Zhu, Z.; Wang, Y.; Anzalone, A.V.; Raguram, A.; Doman, J.L.; et al. Prime genome editing in rice and wheat. Nat. Biotechnol. 2020, 38, 582–585. [Google Scholar] [CrossRef]
  184. Saeglitz, C.; Bartsch, D. Regulatory and associated political issues with respect to Bt transgenic maize in the European Union. J. Invertebr. Pathol. 2003, 83, 107–109. [Google Scholar] [CrossRef]
  185. Jones, H.D. Challenging regulations: Managing risks in crop biotechnology. Food Energy Secur. 2015, 4, 87–91. [Google Scholar] [CrossRef] [Green Version]
  186. Eckerstorfer, M.F.; Grabowski, M.; Lener, M.; Engelhard, M.; Simon, S.; Dolezel, M.; Heissenberger, A.; Luthi, C. Biosafety of genome editing applications in plant breeding: Considerations for a focused case-specific risk assessment in the EU. BioTech 2021, 10, 10. [Google Scholar] [CrossRef]
  187. Agapito-Tenfen, S.Z.; Okoli, A.S.; Bernstein, M.J.; Wikmark, O.G.; Myhr, A.I. Revisiting risk governance of GM plants: The need to consider new and emerging gene-editing techniques. Front. Plant Sci. 2018, 9, 1874. [Google Scholar] [CrossRef] [Green Version]
  188. Busch, G.; Ryan, E.; von Keyserlingk, M.A.G.; Weary, D.M. Citizen views on genome editing: Effects of species and purpose. Agric. Human Values 2022, 39, 151–164. [Google Scholar] [CrossRef]
  189. Pausch, P.; Al-Shayeb, B.; Bisom-Rapp, E.; Tsuchida, C.A.; Li, Z.; Cress, B.F.; Knott, G.J.; Jacobsen, S.E.; Banfield, J.F.; Doudna, J.A. CRISPR-casphi from huge phages is a hypercompact genome editor. Science 2020, 369, 333–337. [Google Scholar] [CrossRef] [PubMed]
  190. Calvache, C.; Vazquez-Vilar, M.; Selma, S.; Uranga, M.; Fernandez-Del-Carmen, A.; Daros, J.A.; Orzaez, D. Strong and tunable anti-CRISPR/Cas activities in plants. Plant Biotechnol. J. 2022, 20, 399–408. [Google Scholar] [CrossRef] [PubMed]
  191. Garrett, R.; Niles, M.T.; Gil, J.D.; Gaudin, A.; Chaplin-Kramer, R.; Assmann, A.; Assmann, T.S.; Brewer, K.; de Faccio Carvalho, P.C.; Cortner, O.J.A.S. Social and ecological analysis of commercial integrated crop livestock systems: Current knowledge and remaining uncertainty. Agric. Syst. 2017, 155, 136–146. [Google Scholar] [CrossRef]
Figure 1. Different types of sequence-specific nucleases and types of editing. (A) Meganucleases, zinc finger nuclease (ZFN), transcription activator-like effector nuclease (TALEN), and CRISPR/Cas9 induces double-stranded breaks, which were corrected by non-homologous end-joining (NHEJ) and homologous recombination (HR). (B) Schematic diagram of target insertion, target deletion, and chromosomal arrangement through genome editing technologies. InDel, insertion-deletion.
Figure 1. Different types of sequence-specific nucleases and types of editing. (A) Meganucleases, zinc finger nuclease (ZFN), transcription activator-like effector nuclease (TALEN), and CRISPR/Cas9 induces double-stranded breaks, which were corrected by non-homologous end-joining (NHEJ) and homologous recombination (HR). (B) Schematic diagram of target insertion, target deletion, and chromosomal arrangement through genome editing technologies. InDel, insertion-deletion.
Plants 11 02625 g001
Figure 2. General procedures in plant transformation, delivery methods of CRISPR cargo, and transgene-free mutant development. (A) Major steps in plant genome editing. Once transformation vectors are designed and constructed, their activity may be validated with protoplasts before being delivered into the host plant. Protoplast transformation can also be used directly to produce transformed cells and eventually plants as described in panel D. The general procedure of transformation also usually followed by a selection process to select resistant cells and seedlings, and finally a sequencing process to confirm presence of transformed mutants. (B) Plant genome editing via Agrobacterium-mediated delivery of CRISPR DNA. Agrobacterium containing the vectors are transfected into plant cells in the form of calli, embryos, or leaf explants, followed by the selection process to produce genome-edited plants (C) Conventional and transient expression approaches for particle bombardment-mediated genome editing via CRISPR DNA, RNA, or RNP delivery. Transformation vectors-coated gold particles are bombarded into plant cells followed by the selection process (D) Protoplast transformation with CRISPR DNA, RNA, or RNP. Transformation vectors, protoplasts, PEG, and Ca2+ ions are mixed before further selection processes to isolate transformed calli, seedlings, and finally genome-edited plants. (E) Two ways to obtain transgene-free mutants. Using the conventional method, a selection agent is used to select resistant calli and transgenic plants. Transformation vectors can be segregated out from the mutant genomes via selfing or crossing. Using the transient method, no selection agent is needed to segregate out the transformation vectors to produce transgene-free mutants. [RNP, ribonucleotide protein].
Figure 2. General procedures in plant transformation, delivery methods of CRISPR cargo, and transgene-free mutant development. (A) Major steps in plant genome editing. Once transformation vectors are designed and constructed, their activity may be validated with protoplasts before being delivered into the host plant. Protoplast transformation can also be used directly to produce transformed cells and eventually plants as described in panel D. The general procedure of transformation also usually followed by a selection process to select resistant cells and seedlings, and finally a sequencing process to confirm presence of transformed mutants. (B) Plant genome editing via Agrobacterium-mediated delivery of CRISPR DNA. Agrobacterium containing the vectors are transfected into plant cells in the form of calli, embryos, or leaf explants, followed by the selection process to produce genome-edited plants (C) Conventional and transient expression approaches for particle bombardment-mediated genome editing via CRISPR DNA, RNA, or RNP delivery. Transformation vectors-coated gold particles are bombarded into plant cells followed by the selection process (D) Protoplast transformation with CRISPR DNA, RNA, or RNP. Transformation vectors, protoplasts, PEG, and Ca2+ ions are mixed before further selection processes to isolate transformed calli, seedlings, and finally genome-edited plants. (E) Two ways to obtain transgene-free mutants. Using the conventional method, a selection agent is used to select resistant calli and transgenic plants. Transformation vectors can be segregated out from the mutant genomes via selfing or crossing. Using the transient method, no selection agent is needed to segregate out the transformation vectors to produce transgene-free mutants. [RNP, ribonucleotide protein].
Plants 11 02625 g002
Figure 3. Co-occurrence network of 50 most used keywords in CRISPR-related plant research from 2012 to 2021. (A) Network visualization of the keywords based on total link strength. Green, yellow, red, and blue nodes represent four different clusters of keywords identified. A minimum strength of 40 was set for the lines to appear between the nodes. The relatedness of the keywords depends on the number of articles in which they occur together, which is indicated by the size of the nodes/keywords, and the length/thickness of the lines between the nodes. The bigger the nodes/keywords, the larger the weight of the nodes/keywords. The shorter and thicker the lines between the nodes, the more frequently they appear together in the publications. (B) Density visualization of the keywords based on occurrences. The density of a keyword depends on the number of keywords around the node. Keywords in the yellow areas indicate a more frequent occurrence in the publications while green areas indicate a less frequent appearance.
Figure 3. Co-occurrence network of 50 most used keywords in CRISPR-related plant research from 2012 to 2021. (A) Network visualization of the keywords based on total link strength. Green, yellow, red, and blue nodes represent four different clusters of keywords identified. A minimum strength of 40 was set for the lines to appear between the nodes. The relatedness of the keywords depends on the number of articles in which they occur together, which is indicated by the size of the nodes/keywords, and the length/thickness of the lines between the nodes. The bigger the nodes/keywords, the larger the weight of the nodes/keywords. The shorter and thicker the lines between the nodes, the more frequently they appear together in the publications. (B) Density visualization of the keywords based on occurrences. The density of a keyword depends on the number of keywords around the node. Keywords in the yellow areas indicate a more frequent occurrence in the publications while green areas indicate a less frequent appearance.
Plants 11 02625 g003
Table 1. Examples of CRISPR/Cas9 applications for crop improvement.
Table 1. Examples of CRISPR/Cas9 applications for crop improvement.
ImprovementTraitCropsgRNA Target AreaType of EditingTarget AreaResultReferences
Abiotic stress resistanceDroughtChickpeacDNAFrameshift deletionCoumarate ligase (4CL) and Reveille 7 (RVE7)Enhanced tolerance[85]
ColdRicecDNAInDel mutationOsMYB30Improved tolerance[86]
HerbicideMaizecDNA Base editingZmALS1, ZmALS2Plants with Sulfonylurea herbicide-resistant [87]
SalinityTomatoDBD domain of cDNA49-bp deletionSlARF4Enhanced salinity tolerance[88]
Heavy metalsRicecDNADownregulationOsNramp5Decreased cadmium accumulation[89]
HeatTomatocDNA1-bp insertion
4-bp deletion
SlMAPK3Enhanced heat tolerance[90]
Biotic stress resistanceViral diseaseBarleyCoding sequenceBase editingMP, CP, Rep/Rep, IR/Virus genomeResistant plants[91]
Fungal diseaseRiceGenome80-bp insertALB1, RSY1/ Fungal geneImproved resistance to rice blast[92]
Bacterial diseaseTomatoJAS domain C-terminalDeletionSIDMR6-1/Host S-geneResistant plants[93]
Insect pestSoybeanCoding region1-bp and 33-bp deletionGmUGTEnhanced resistance to Helicoverpa armigera and Spodoptera litura[94]
Plant/crop qualityCrop growthRicecDNAFrameshiftPYL1–PYL6 and PYL12(gp-1), PYL7–PYL11 and PYL13(gp-2)Improved plant growth and grain productivity[95]
Crop yieldWheatcDNA10-bp deletionTaCKX2-1, TaGLW7, TaGW2, and TaGW8Improved grain yield[96]
Crop nutritionRiceGenomic Safe Harbor5.2kb insertion5.2 kb carotenoid cassette insertionIncreased β-carotene content[97]
Grain sizeRicecDNAInDel mutationOsGS3Increased grain size[98]
Grain numberRicecDNAInDel mutationOsGn1aIncreased grain number[98,99]
Fruit sizeTomatoPromoter85-bp deletionSlENOEnhanced fruit size[100]
Table 2. A list of 50 most frequently occurring keywords in CRISPR-related plant research publications from 2012 to 2021. The ranking is based on the number of occurrences in the publications. Total link strength indicates the number of publications in which two keywords occur together.
Table 2. A list of 50 most frequently occurring keywords in CRISPR-related plant research publications from 2012 to 2021. The ranking is based on the number of occurrences in the publications. Total link strength indicates the number of publications in which two keywords occur together.
RankKeywordOccurrencesTotal Link StrengthRankKeywordOccurrencesTotal Link Strength
1crispr2386626026chloroplast146384
2CRISPR/Cas9821187327plasmid146499
3plant protein769238528crispr/cas144267
4arabidopsis673211129transgene141542
5human535162930protoplast140494
6crop531145031soybean134453
7rice525146232enzyme128437
8gene514167733flower124427
9plant512159734quantitative trait locus123405
10animal409128735transcription activator like effector nuclease117494
11plant disease33599136chromosome110376
12transcription factor292100837microrna110387
13agrobacterium282102638mitochondrion108285
14protein26293839double stranded dna break105440
15tomato23977240plant cell105375
16wheat22469041bacterial protein99414
17plant leaf22081442plant virus96310
18maize21374243fungus91314
19allele19570844drought76211
20esterase19339845intron75156
21tobacco19269946host pathogen interaction72267
22bacterium18162547cas70292
23site-directed mutagenesis17063048mouse60215
24endonuclease15568749fatty acid54159
25plant root15249850recombinant protein53192
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hamdan, M.F.; Karlson, C.K.S.; Teoh, E.Y.; Lau, S.-E.; Tan, B.C. Genome Editing for Sustainable Crop Improvement and Mitigation of Biotic and Abiotic Stresses. Plants 2022, 11, 2625. https://doi.org/10.3390/plants11192625

AMA Style

Hamdan MF, Karlson CKS, Teoh EY, Lau S-E, Tan BC. Genome Editing for Sustainable Crop Improvement and Mitigation of Biotic and Abiotic Stresses. Plants. 2022; 11(19):2625. https://doi.org/10.3390/plants11192625

Chicago/Turabian Style

Hamdan, Mohd Fadhli, Chou Khai Soong Karlson, Ee Yang Teoh, Su-Ee Lau, and Boon Chin Tan. 2022. "Genome Editing for Sustainable Crop Improvement and Mitigation of Biotic and Abiotic Stresses" Plants 11, no. 19: 2625. https://doi.org/10.3390/plants11192625

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