Next Article in Journal
Hemin with Peroxidase Activity Can Inhibit the Oxidative Damage Induced by Ultraviolet A
Next Article in Special Issue
Juglone from Walnut Produces Cardioprotective Effects against Isoproterenol-Induced Myocardial Injury in SD Rats
Previous Article in Journal
The Pathological Links between Adiposity and the Carpal Tunnel Syndrome
Previous Article in Special Issue
Exogenous Zeaxanthin Alleviates Low Temperature Combined with Low Light Induced Photosynthesis Inhibition and Oxidative Stress in Pepper (Capsicum annuum L.) Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CRISPR/Cas9 Technique for Temperature, Drought, and Salinity Stress Responses

1
Graduate School of Life and Environmental Sciences, University of Tsukuba, Tsukuba 305-8572, Japan
2
Tsukuba-Plant Innovation Research Center, University of Tsukuba, Tsukuba 305-8572, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Curr. Issues Mol. Biol. 2022, 44(6), 2664-2682; https://doi.org/10.3390/cimb44060182
Submission received: 12 May 2022 / Revised: 2 June 2022 / Accepted: 6 June 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Advanced Research in Plant Metabolomics)

Abstract

:
Global warming and climate change have severely affected plant growth and food production. Therefore, minimizing these effects is required for sustainable crop yields. Understanding the molecular mechanisms in response to abiotic stresses and improving agricultural traits to make crops tolerant to abiotic stresses have been going on unceasingly. To generate desirable varieties of crops, traditional and molecular breeding techniques have been tried, but both approaches are time-consuming. Clustered regularly interspaced short palindromic repeat/Cas9 (CRISPR/Cas9) and transcription activator-like effector nucleases (TALENs) are genome-editing technologies that have recently attracted the attention of plant breeders for genetic modification. These technologies are powerful tools in the basic and applied sciences for understanding gene function, as well as in the field of crop breeding. In this review, we focus on the application of genome-editing systems in plants to understand gene function in response to abiotic stresses and to improve tolerance to abiotic stresses, such as temperature, drought, and salinity stresses.

1. Introduction

According to the World Meteorological Organization (WMO), atmospheric greenhouse gas concentrations and the associated accumulated heat have hit record highs. In the last ten years, conflicts, global climate changes, and economic shocks have increased in frequency and intensity. The combined consequences of these threats, exacerbated by the COVID-19 pandemic, have resulted in about a 20% rise in hunger, from 135 million people in 2020 to 161 million by September 2021 [1].
Extreme weather events affect crop yield and food production, which exert inimical effects on the agronomic industry and, thus, increase the demand for agricultural products [2]. Increased concentrations of atmospheric CO2 are expected to improve water use efficiency and photosynthetic rates in crops and increased cumulative temperatures may prolong the growing season for crops and reduce growth cycles, particularly in mid-latitudes location [3]. However, it has been strongly proven that the negative impacts of climate change on agricultural and terrestrial food production are more common than the positive ones. Food security is facing large global and regional risks owing to climate change and increasing food demand. Thus, food production is an important aspect of food security [4]. Major abiotic stressors, including drought, extreme temperatures, and salinity, cause extensive losses in agricultural production [5]. It is estimated that adverse environmental stresses reduce crop yield from 50 to 70%. [6]. The process of plant responses to abiotic stresses is complicated due to the interactions between stressors and diverse molecular, biochemical, and physiological aspects [7]. The development of plant tolerance to abiotic stress is considered a promising approach for increasing crop yields. The knowledge and progress of the structural and functional components of crop genomes have been aided by modern genetics and breeding approaches. Advances in genomics and genome editing have broadened the chances of crop improvement [8]. Since modifying a specific genetic locus is an efficient method for improving breeding, genome-editing technology is increasingly being used in plants.
Because of global population growth, food demand has increased, and sustainable agriculture should be promoted to address the issues of climate change. It is necessary to increase crop varieties to cope with abiotic stresses. Traditional selection or breeding methods, such as marker-assisted selection, are ways to produce crop varieties. In addition, genome editing technologies, in particular the clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9) system, are biotechnological techniques for precise genetic modification of crops and can increase varieties with a reduction of time consumption for breeding. In this review, we describe the CRISPR/Cas9 system and highlight the use of the technique to produce plant and crop varieties for temperature, drought, and salinity stress responses. Valuable information concerning the tolerance and sensitive to these stresses is described. The information about tolerance to these stresses can be used for future molecular-based breeding approaches to produce these stress-tolerant plants, and the information about sensitivity can also be useful to understand the molecular mechanisms of the responses to temperature, drought, and salinity stresses. We list the information of several kinds of plants, including Zea mays (maize), Oryza sativa (rice), Arabidopsis thaliana (Arabidopsis), Solanum lycopersicum (tomato), Glycine max (soybean), Triticum aestivum (wheat), and Solanum tuberosum (potato). The list provides important information about targets for genome editing in breeding new abiotic stress-tolerant varieties or the application of this information to other plants.

2. Genome-Editing Technology

Genome editing, a revolutionary and accurate genetic-engineering technology that can modify specific target genes of the organism genome, is increasingly used in many fields including plant science and crop breeding [9]. Genome-editing technology plays a role in the characterization of gene function and crop improvement [10]. There are three main types of genome-editing technology: zinc-finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and the CRISPR/Cas system [11].
A wide spectrum of genetic alternation has been performed by ZFNs and TALENs via inducing DNA double-strand breaks, which stimulate error-prone non-homologous end joining (NHEJ) or homology-directed repair (HDR) at specific genomic locations [12]. A class of targeting reagents, ZFNs, consists of zinc-finger-based DNA-recognition modules and the DNA cleavage domain of the FokI restriction enzyme [13,14]. Each individual zinc finger recognizes and attaches to a nucleotide triplet and then assembles into groups to bind to particular DNA sequences. Because ZFNs with a high sequence affinity are complex to design and their off-target efficiency is high [15], it took 9 years from the development of ZFNs to the success of the first ZFN-based plant genome editing [16]. The transcriptional activator-like effector (TALE) repeats and the FokI restriction enzyme are fused together to form TALENs (Figure 1A). The central domain of TALE consists of a repeating unit composed of approximately 34 amino acids. The specificity of TALEs depends mainly on repeat variable diresidues (RVDs) and hypervariable amino acids at the 12th and 13th positions. Four different amino acid RVDs (NI, HD, NG/HG, and NN) were used to identify adenine (A), cytosine (C), thymine (T), and guanine (G)/adenine (A), respectively. The DNA identifiers provide the intimate connection between the network of amino acid repeats and the nucleotide sequence of the genome; hence, TALENs can be engineered with desired sequence properties [17].
Currently, a popular and powerful gene-editing tool is the CRISPR/Cas9 system, because of its simplicity and usability. Cas enzymes are divided into two classes (Classes I and II) depending on the architecture of interference effector modules and six major types (types I–VI); 33 subtypes exist based on the signature cas gene and distinct mechanism of targeting [18]. As part of the bacterial adaptive immune system, CRISPR sequences can recognize exogenous viral DNA and directly cleave Cas proteins [19]. Among these subtypes, the CRISPR/Cas9 system has been extensively studied. The Cas9 protein and synthetic single-guide RNA (sgRNA), which is generated by fusing CRISPR-RNA (crRNA) and transactivation RNA (tracrRNA), are the primary components of the CRISPR/Cas9 system. The site-specific DNA double-strand breaks (DSBs) are introduced by sgRNA-guided Cas9 protein, thus triggering DNA repair mechanisms (Figure 1B). The target sequence located upstream of the protospacer-associated motif (PAM), NGG, in the case of CRISPR/Cas9, was designed. The system has been widely used in plants for characterization of gene function and precision plant breeding [20,21,22].
Although the CRISPR/Cas9 system is effective, other Cas proteins have been identified and applied in genome-editing technology. Cas12a/Cpf1 is the most notable Cas protein in addition to Cas9. It recognizes different PAM sequences and produces sticky ends, rather than blunt ends, which facilitates HDR repair to produce more precise editing [23]. Among the Cas enzyme subtypes, some have been applied to plants, such as Cas3 [24], Cas10 [24], Cas12 [23], Cas12b [25], Cas13a [26,27], and CRISPR/CasΦ [28]. These different enzymes may increase the range of options available for different plant species. SpCas9 derived from Streptococcus pyogenes has been widely used. Compared to SpCas9, SaCas9 derived from Staphylococcus aureus is smaller and more conducive to intracellular delivery [29]. In addition, researchers have identified many Cas proteins targeting different PAM sequences, such as StCas9 (from Streptococcus thermophilus), NmCas9 (from Neisseria meningitidis), FnCpf1 (from Francisella novicida), and so on [30].
Base-editing technology has become another option for effective and accurate genome editing. The base editor enables precise substitution of individual nucleotides. Fusion of dead (dCas9) or nickase Cas9 (nCas9) with cytidine base editing (CBE) and adenine base editing (ABE) can achieve base editing from C–G to T–A and from A–T to G–C, respectively [31,32]. Because single-nucleotide mutations are important sources of superior crop traits [33,34], base editors can improve crop quality.
In polyploids, many genes consist of several copies, named homologs, which have the same role in determining certain plant features. These homologs must be altered concurrently in order to produce recessive mutants. Furthermore, members of the gene family usually share comparable structures and activities, contributing to genetic robustness. Therefore, it will be more efficient to mutate two or more paralogs to show a phenotypic effect rather than knocking out one gene. The multiplex genome-editing technique is a worthwhile technique to genetically alter several gene functions [35]. This technique has demonstrated its great superiority in editing multiple target sites simultaneously. Recent research using multiplex genome editing in plants has been predominantly carried out by the CRISPR/Cas system, because an sgRNA can target several homologous alleles. If a conserved sequence exists in several genes, an sgRNA can target them all or multiple gRNAs can also be employed to target genes that contain no conserved sequence [36].
Prime editing is an emerging genome-editing technology to achieve precise editing and consists of two parts: one is the effector protein formed by the fusion of nCas9 and the engineered reverse transcriptase (RT); the other is the prime editing guide RNA (pegRNA) including sgRNA, primer binding site (PBS), and RT template with edit. A pegRNA is a special gRNA that guides the edited protein to the target site and contains the edited template sequence. The Cas9-RT fusion protein precisely cuts a DNA strand under the guidance of pegRNA and then synthesizes DNA with the correct sequence according to the template. The DNA repair machinery in the cell automatically integrates this newly synthesized sequence into the genome [37]. Compared with the CRISPR/Cas9 system, the restriction of PAM sequences on prime editing is greatly reduced. For prime editing, the distance from the PAM to the editing site can exceed 30 base pairs (bp). Prime editing can achieve all 12 base substitutions, while the base editor can only perform four types of substitutions [38]. Even though CRISPR/Cas9-mediated genome editing can be used for HDR after double-strand breaks, the efficiency of HDR production in plants is very low. In contrast, the advent of pregRNAs has greatly improved editing efficiency. Because 33 of 62 transgenic lines harbor the substation of serine 621 to isoleucine (S621I) in the ZmALS1 and/or ZmALS2 genes, the editing efficiency of prime editing in maize is as high as 53.2% (33/62) [39]. As an emerging genome-editing method, prime editing still needs more exploration. However, its own characteristics of high efficiency, precision, and site-directed mutation make it have infinite prospects and wide applications in future gene-editing projects.

3. Abiotic Stress Responses Revealed by Genome Editing

One of the innovative techniques, the CRISPR/Cas9 system precisely introduces mutation(s) in the target gene(s), resulting in enhancement of tolerance to abiotic stresses or the discovery of the mechanism of signaling. Table 1 summarizes the application of CRISPR/Cas9 to identify the genes for response to abiotic stresses, such as temperature, drought, and salinity stresses.

3.1. Temperature Stress Responses

Heat and cold stresses can have dramatic impacts on agriculture the and ecology of plants. Leaf photosynthesis, biomass accumulation, and grain yield are affected by cold stress [79].
Proline-rich proteins function in a variety of physiological and biochemical processes for plant growth and stress responses. The knockout of OsPRP1, which encodes a proline-rich protein (PRP) in rice generated by CRISPR/Cas9, enhances cold sensitivity in rice [40]. The CRISPR/Cas9-mediated knockout mutant of OsMYB30, which has been characterized as a cold-responsive R2R3-type MYB gene in rice, exhibits better cold tolerance than wild-type rice [41].
The three tandemly arranged CCAAT-binding factor (CBF) genes—CBF1, CBF2, and CBF3—have been verified to be involved in cold acclimation using the CRISPR/Cas9-mediated Arabidopsis Atcbf single, double, and triple mutants. Compared with wild-type Arabidopsis plants, cold-acclimated Atcbf triple mutants exhibit extremely sensitive phenotype to freezing temperature. Expression of CBF genes is rapidly increased when plants are exposed to cold temperatures. The CBF proteins enhances the transcription of downstream cold-responsive (COR) genes to increase the freezing tolerance of plants [42]. The freezing sensitivity ranking is triple cbf mutant > cbf1 cbf3 double mutant > cbf3 mutant [43]. Among the tomato CBF gene family, SlCBF1 is the only cold-inducible gene [80]. Both salicylic acid (SA)- and hydrogen peroxide (H2O2)− induced cold tolerance are achieved by increased SlCBF1 expression in tomato [44]. The Slcbf1 mutants generated by the CRISPR/Cas9 system exhibited greater electrolyte leakage and malondialdehyde (MDA) levels than wild-type plants, indicating that knockout of SlCBF1 can increase cold-stress-induced membrane damage [45].
Many of the cis-regulatory elements in the rice OsAnn5 promoter region are common promoter elements. Some elements are unique to OsAnn5, including the dehydration-responsive element (DRE) core (a cis-acting element involved in CBF-mediated cold responsiveness) and MYB recognition sites. In the region between the start codon ATG and −2082 bp of OsAnn5, plant hormone-regulatory elements, as well as light-responsive elements are included. This implies that several transcription factors regulate OsAnn5 expression in rice. Compared with wild-type rice, CRISPR/Cas9-mediated elimination of OsAnn5 function significantly increases survival rates at the seedling stage under cold stress in rice, demonstrating that OsAnn5 is a positive factor that regulates cold stress tolerance at the seedling stage [46]. Plant annexins are calcium-ion (Ca2+)-dependent phospholipid-binding proteins that are implicated in the regulation of plant development and protection from environmental stresses. OsAnn3-knockoout mutants generated by CRISPR/Cas9 exhibit an increase in relative electrical conductivity (REC) and a decrease in survival rate compared with wild-type plants, which proves that OsAnn3 plays a role in cold tolerance in rice [47].
Heat stress adversely affects biological activities, such as chlorophyll content, turgor, photosynthetic rate, carbon assimilation, acquisition of nutrient, cellular metabolism, and gaseous exchange at the leaf surface. Together, these undesirable influences cause a reduction in crop yield [81,82]. The deletion mutants of heat-stress sensitive albino 1 (OsHSA1) generated by CRISPR/Cas9 exhibit greater sensitivity to heat than wild-type rice plants [48]. OsHSA1 encodes fructokinase-like protein2 and functions in chloroplast development and chloroplast protection under heat stress during the early and late stages in rice, respectively. Tomato Slcpk28 (calcium-dependent protein kinase28) mutants produced by CRISPR/Cas9 exhibit increased heat-stress-induced accumulation of reactive oxygen species (ROS) and protein oxidation level and decreased ascorbate peroxidase (APX) activity and other antioxidant enzymes, such as catalase and glutathione reductase, leading to heat sensitivity [49]. CRISPR/Cas9-mediated knockout of SlMAPK3 exhibited better heat tolerance by reducing ROS accumulation and upregulating the expression of genes encoding heat stress transcription factors (HSFs) and heat shock proteins (HSPs) [50]. The tomato BRASSINAZOLE RESISTANT 1 (SlBZR1) gene acts as a critical regulator of the brassinosteroid (BR) response. BRs are a group of steroid plant hormones that can enhance tolerance to a range of abiotic stressors [83]. Heat-stress-induced damage was exacerbated in the bzr1 mutants, and BR-induced heat stress tolerance was lost through RESPIRATORY BURST OXIDASE HOMOLOG1 (RBOH1)-dependent ROS signaling, which is regulated by FERONIA (FER) homologs [51]. Under heat stress conditions, knockout of OsNAC006 using the CRISPR/Cas9 system displayed an obvious increase in superoxide radical (O2) and H2O2 levels, as well as a decrease in chlorophyll content and the activities of antioxidant enzymes, demonstrating that Osnac006 may function in heat tolerance by reducing the antioxidant response, which is triggered in response to oxidative stress and mediates photosynthesis [52]. Although several reports provide molecular-based modification, which confers plant tolerance to temperature stresses, more research to identify the functional mechanism is required. To address the global warming issue, heat stress tolerance varieties and identification of the target to increase heat tolerance are necessary. One of the targets could be PHTOCHROME (PHY). PHYB has been identified as a thermosensor [84,85], and the phy mutant exhibited high tolerance to heat stress in Arabidopsis [86] and tomato [87]. This mutant information is important for the definition of targets of genome editing.
Heat stress is commonly characterized as a temperature rise that exceeds a threshold level for an extended period of time, causing irreversible damage to plant growth and development. Generally, when the ambient temperature is 10–15 °C higher than the optimum range of temperature for crop cultivation, such conditions are defined as heat shock or heat stress [88]. High temperatures can cause high initial rate respiration in pollen, which leads to the exhaustion of endogenous respiratory substrates, subsequent aging, and a loss of mitochondrial activity, resulting in the abortion of the pollen and reduction of the fruit set [89,90]. Parthenocarpy, the development of fruit in the absence of pollination or fertilization, is considered a valuable goal during seedless fruit development because of its fertilization-independence, consumers’ preference for seedless over seeded fruits, and higher fruit quality [91,92]. After screening an ethyl-methanesulfonate (EMS)-mutagenized tomato population for yield under heat stress, a mutant capable of generating high-quality seedless (parthenocarpic) fruit was isolated. Following next-generation sequencing, marker-assisted mapping and CRISPR/Cas9 gene knockout research discovered that the parthenocarpic phenotype was caused by a mutation in the tomato SlAGAMOUS-LIKE 6 (SlAGL6) gene encoding MADS-box. Mutations in SlAGL6 increase tomato yield under heat stress. The tomato agl6 mutants exhibited facultative parthenocarpy without any pleiotropic effect and developed seedless fruits, which were comparable in both weight and shape to wild-type seeded fruits [53]. Aux/IAA9 (IAA9) is involved in tomato fruit development and represses fruit initiation without fertilization. CRISPR/Cas9-mediated IAA9 (SlIAA9) mutant plants exhibited simple leaves instead of wild-type compound leaves, and fruit development of the Sliaa9 mutant was stimulated before fertilization, leading to parthenocarpy. These mutations are heritable in subsequent generations [54]. DELLA is a negative regulator of gibberellin signaling. Loss-of-function mutations in SlDELLA, also called PROCERA, exhibit parthenocarpy in tomatoes [93]. The base-edited Sldella mutant with the Target-AID system exhibited high gibberellin sensitivity and a parthenocarpic phenotype [55]. These results suggest that genome-editing techniques enhance parthenocarpy in tomatoes. Parthenocarpy is an important trait to increase the yield of tomatoes in summertime without the application of auxin. Because tomatoes are a member of Solanaceae, it is possible to knockout the genes by the genome-editing technique in other Solanaceae, such as peppers and eggplants.

3.2. Drought Stress Responses

Drought is one of the abiotic stresses on plants that seriously reduces crop productivity. The tomato SlLBD40 gene, which belongs to subfamily II of the LATERAL ORGAN BOUNDARIES DOMAIN (LBD) family, is highly expressed in roots and fruits. The average water loss rate of wild-type tomato plants was significantly higher than Sllbd40 knockout mutants generated using CRISPR/Cas9 [56]. Auxin response factors (ARFs) are key proteins for various physiological processes, such as leaf expansion, lateral root development, and fruit development, in plants. Knocking out SlARF4 using CRISPR/Cas9 led to lower water loss than in wild-type tomato plants. The leaves of Slarf4 plants were able to stand upright again after 24 h of re-watering when they were wilted, but the wild-type plants were not [57]. AITRs are a family of novel transcription factors that regulate plant responses to abscisic acid (ABA) and abiotic stress in Arabidopsis. Single, double, and triple aitr mutants exhibited increased drought tolerance. Arabidopsis aitr mutants with all six AITR genes being knocked out by CRISPR/Cas9 exhibited enhanced drought and salt tolerance and reduced sensitivity to ABA, but the plant growth and development of the sextuple mutants were not affected [3]. CRISPR/Cas9-mediated mutation in the open stomata 2 (OST2) gene, which encodes a plasma membrane H+ ATPase responsible for stomatal response [94], exhibited a high degree of stomatal closure with a low level of water loss, leading to enhanced drought tolerance in Arabidopsis [59]. Abscisic-acid (ABA)-responsive element binding protein 1/ABRE binding factor (AREB1/ABF2) is an important positive regulator for drought stress response. To activate the endogenous promoter of AREB1, the CRISPR/dCas9HAT system was employed. Arabidopsis histone acetyltransferase (HAT) enhances chromatin relaxation and promotes gene expression. HAT fused with dCas9 increased the promoter activity, and the expression levels of AREB1 were considerably upregulated by 2-fold. The plants generated in this study exhibited better survival rates after drought stress [60]. Maize ARGOS8 negatively regulates ethylene response. Translocation of the GOS2 promoter to the ARGOS8 locus using CRISPR/Cas9 and ARGOS8 variants resulted in elevated levels of ARGOS8 transcripts. These variants were compared with a wild-type hybrid under drought stress conditions, and their yields did not decrease significantly. These results indicate that ARGOS8 variants improve maize grain yield under field drought stress conditions [61]. OsERA1 acts as a negative regulator of responses to drought stress in rice. Osera1 mutant seedlings, which were created using CRISPR/Cas9, were subjected to water-deficit stress. The relative stomatal conductance rates of the mutant plants were significantly lower than those of the wild-type plants. The Osera1 mutant exhibits increased sensitivity to ABA [62]. Leaf morphology affects plant tolerance to drought, and genes for the curled leaf phenotype by controlling the quantity, size, and arrangement of bulliform cells (BCs) were influenced. The rolled leaf phenotype was observed by the CRISPR/Cas9-based mutagenesis of Semi-rolled leaf1,2 (SRL1 and SRL2) genes. Mutant plants had lower MDA levels and higher content of abscisic acid (ABA) than wild-type plants under drought stress. This indicates that the mutant plants are more drought-tolerant than the wild type [63].
In contrast, some genes have been edited to make plants drought sensitive. These mutants are also useful to understand more precisely the mechanism of drought stress responses. A master regulator, NPR1, is involved in plant defense responses to pathogens. Interestingly, tomato Slnpr1 mutants created by CRISPR/Cas9 exhibited reduced drought tolerance, and the survival rate of Slnpr1 mutants was significantly lower than that of wild-type plants under drought conditions [64]. Mitogen-activated protein kinases (MAPKs) are key signaling molecules. Compared with wild-type plants, slmapk3 mutants created using the CRISPR/Cas9 system exhibited more severe wilting and membrane damage under drought stress in tomato [65]. The MYB-related gene GmMYB118 improved tolerance to drought stress by reducing the content of ROS and MDA. Using CRISPR/Cas9 to knock out GmMYB118, mutant plants exhibited reduced drought tolerance [66]. OsPUB67, which encodes a U-box E3 ubiquitin ligase in rice, was drastically induced by drought stresses. A drought-sensitive phenotype was observed in the CRISPR/Cas9 knockout plants. At the seedling stage, the drought stress response of the overexpressing (OE) lines and mutant lines was investigated. The survival rates of the OE lines were significantly greater after re-watering than those of the wild type and mutants. These results suggest that OsPUB67 has a positive role in drought stress tolerance [67]. Osmotic stress-/ABA-activated protein kinase 2 (SAPK2) is the primary mediator of ABA signaling. Loss-of-function mutants were produced using the CRISPR/Cas9 system. Compared with the wild type, sapk2 mutants are mostly insensitive to ABA, with relatively low survival rates under drought stress [68].
As indicated, several mutations generated by CRISPR/Cas9 enhanced drought tolerance. It is possible that stacking of mutations by crossing or by multiplexing will show additive effects, enabling the production of plants highly tolerant to drought stress.

3.3. Salinity Stress Responses

Salinity is a severe threat to crop yield, because most agricultural plants cannot survive in high-salt environments [95]. Research on salinity tolerance is becoming increasingly important, and enhancement of salt tolerance has become an important breeding goal. The OsRR22 gene, which encodes a 696-amino-acid B-type response regulator transcription factor, is involved in both cytokinin signal transduction and metabolism. The Cas9-mediated OsRR22 mutant lines grew better than the wild type under a 0.75% sodium chloride (NaCl) nutrition solution [69]. OsVDE, a lipocalin-like protein in chloroplasts, facilitates ABA biosynthesis and negatively regulates salt-stress tolerance in rice seedlings. The Osvde mutant generated by the CRISPR/Cas9 system exhibited greater stomatal closure and higher ABA content than the wild type, resulting in the reduction of water loss from the transpiration process [70]. The drought and salt tolerance (DST) gene was selected for CRISPR/Cas9-mediated genome editing in indica rice cv. MTU1010, and different mutant alleles of the DST gene were successfully generated. dstΔ184–305 were selected for phenotypic analysis. Under osmotic and salt stress, > 65% of dstΔ184–305 mutant seedlings survived. Furthermore, dstΔ184–305 mutants exhibited higher chlorophyll retention [71]. A total of 158 NAC transcription factors have been identified in rice that are involved in multiple abiotic stresses. To determine the particular role of OsNAC041 in abiotic stress response, targeted mutagenesis of the OsNAC041 locus was performed using the CRISPR/Cas9 system. Seed germination and subsequent growth of the Osnac041 mutants were suppressed at 7 days when treated with 150 mM NaCl, relative to the wild type. Furthermore, shoots of the Osnac41 mutant seedlings were shorter than those of the wild-type seedlings under salt stress, demonstrating that the mutants were sensitive to salt stress [72].
Hybrid proline-rich proteins (HyPRPs), a subclass of putative plant cell wall glycoproteins, have been shown to be negative regulators of tomato multi-stress responses. Engineering the SlHyPRP1 gene by precisely removing its PRD, 8CM domain, or both resulted in greater survival rates than the wild type in medium containing 150 mM NaCl [73]. Auxin response factors (ARFs) also have important roles in regulating the expression of auxin response genes in tomatoes. Using CRISPR/Cas9 to obtain SlARF4 mutants under NaCl exposure, the loss of shoot fresh weight in the mutants was half that of the wild type [74].
AITRs are a family of transcription factors that regulate plant responses to abiotic stresses. In Arabidopsis, there are six genes encoding AITRs. Knockout of any one of these genes using CRISP/Cas9 showed enhanced tolerance to salt treatment, indicating that the entire family of AITR genes in Arabidopsis leads to enhanced salinity tolerance [57]. The ACQOS locus is a cluster comprising four nucleotide-binding leucine-rich repeats (NLRs) encoding a toll-interleukin1 receptor-nucleotide-binding leucine-rich repeat class protein. Wild-type Arabidopsis ACQOS alleles are salt sensitive. A knockout line of the ACQOS allele has been established using the CRISPR/Cas9 system in Arabidopsis. Chlorophyll measurements suggested that ACQOS silencing significantly affected salt stress tolerance, because the chlorophyll content was significantly reduced in mutants compared to the wild type [75].
Bread wheat (Triticum aestivum L., BBAADD) is a typical allohexaploid species with higher salt tolerance than its tetraploid wheat progenitor (BBAA) [76]. The TaHAG1 overexpressed (TaHAG1-OE) plants were subjected to salt stress together with wild-type plants, and the TaHAG1-OE lines exhibited a less severe phenotype. To further verify the function of TaHAG1 in salt tolerance, Tahag1 mutants were produced by the CRISPR/Cas9 system. Under salt stress, the wild-type and mutant plants exhibited obvious physiological differences, including a decrease in root length and fresh weight, an increase in chlorotic leaves, and greater Na+ content in the mutants. This indicates that TaHAG1 acts as a crucial regulator in strengthening the salt tolerance of hexaploid wheat [76].
Coilin is the major structural protein that controls the formation, composition, and activity of subnuclear Cajal bodies. A fragment of the potato coilin gene encoding the CTD was edited using CRISPR/Cas9 technology. Under salt-stress conditions, the WT exhibited accelerated yellowing, leaf fall, and more severe inhibition of root development, indicating that potato coilin is involved in the plant defense response to salinity [77].
Soybean growth and yield are largely affected by abiotic stresses such as drought, salinity, and extreme temperatures. AITRs are ABA-induced transcriptional repressors. Transgene-free Gmaitr mutants were generated using CRISPR/Cas9 genome editing to target GmAITR genes. Under salt treatment, Gmaitr mutant seeds showed a higher germination rate than the wild type, and the Gmaitr mutant plants were morphologically similar to the wild-type plants in the normal soil field, but grew better in the saline soil field [78]. Interestingly, CRISPR/Cas9-mediated mutation in the AITR gene enhanced salinity stress tolerance in Arabidopsis and soybean. Furthermore, the aitr mutant also exhibited drought tolerance in Arabidopsis [58]. It is plausible that the AITR gene is one of targets for CRISPR/Cas9-mediated mutagenesis to increase tolerance to salinity stress in several plant species.

4. Perspectives

A growing world population and increasingly frequent climate extremes threaten human food security [96]. However, humans have limited means to increase food production, and traditional mutation breeding techniques rely on unpredictable genomic mutations and labor-intensive screening, resulting in inefficiencies [97]. There is an urgent need for rapid, targeted breeding methods to accelerate plant improvement. Genome editing is a useful technology in crop breeding [9]. In contrast to earlier genetic engineering techniques, which involve the random insertion of a foreign gene into the plant genome, genome-editing technology can manipulate the genome of organisms more precisely [98]. Thus, genome-editing technology has facilitated the acquisition of plants with characteristics of interest. However, several challenges remain to be overcome.

4.1. Cost and Regulation Aspects

One of the main limitations of genome-editing techniques is the cost of producing and using zinc fingers and TALEN. Since the discovery and rise of CRISPR/Cas9, the cost, time, and work required to obtain genome-edited plants have significantly decreased [19,99,100]. Gene-edited products have been available on the market since 2012, when an article about CRISPR/Cas9 was published [101]. The first available product was Calyno, a high-oleic soybean oil produced by Calyxt in the U.S. [102]. In addition to genome-edited soybean, herbicide-tolerant canola by Cibus in Canada and γ-aminobutyric-acid (GABA)-rich tomato have been commercially provided by Sanatech Seed in Japan [103]. Furthermore, genome-edited fishes, such as “Madai” red sea bream with an improved feed utilization efficiency and “22-seiki fugu” tiger puffer for faster growth, have reached consumers.
In addition, to facilitate access to gene-editing technology, unlike techniques for obtaining genetically modified organisms (GMOs), genome-editing techniques do not require the insertion of exogenous DNA, making them unsuitable for the regulatory regimes of GMOs [104]. Despite this, the EU and other countries such as New Zealand have adapted their legislation and consider genome-edited plants as GMOs, stigmatizing their commercialization and acceptance [105]. Research has already shown that the public is more receptive to the consumption of CRISPR-derived products than of GM-derived products [106]. Recently, the European Commission started to reconsider the GMO rules for gene-edited crops. It is possible that genome-edited foods will be widely available on the market.

4.2. Genome-Editing Tools Delivery Limitations

Agrobacterium-based tissue culture methods are often used to deliver CRISPR/Cas9 expression cassettes in plants. However, there are still limitations related to tissue culture and plant regeneration. Tissue culture is time consuming and laborious. Some innovative genome-editing delivery methods have recently emerged, such as viral delivery, nanoparticle (NP) delivery, and meristem induction (Table 2). The virus is an efficient DNA-delivery tool that can deliver sgRNAs; however, it is difficult to use viruses to deliver complete CRISPR/Cas9 expression cassettes because of the limited loading capacity of viral capsid proteins [107]. Rhabdovirus and potato virus X can deliver the complete CRISPR/Cas9 expression cassette and can be successfully applied for genome editing in Nicotiana benthamiana [108,109]. NPs can deliver CRISPR/Cas9 expression cassettes to plant protoplasts [110], and cationic lipid NP-based genome editing has been achieved in citrus [111]. However, the size of the NPs is still large, making it difficult to deliver the complete CRISPR/Cas9 expression cassette through the cell wall. The meristem-induction method is a genome-editing method that does not rely on tissue culture. This method introduces both plant growth regulators and CRISPR/Cas9 expression cassettes into plant leaves, thereby inducing meristematic tissues in the leaves. This method achieves relatively high genome-editing efficiency in Nicotiana benthamiana [112]. In addition, transient expression techniques have been used to improve base editing in various plants [113,114,115,116,117]. Exogenous DNA may not integrate into the plant genome, which facilitates the rapid cultivation of transgene-free plants. The in planta particle bombardment (iPB) method is another method to deliver DNA or CRISPR/Cas9-ribonucleoprotein to shoot apical meristem cells [118]. In wheat, 5.2% of plants carried mutant alleles, and mutations in three of these were also inherited in the following generation. Because Cas9 and sgRNA were not detected, the transient expression of CRISPR/Cas9 introduced mutations [119]. This technique has also been successfully applied to other species such as sorghum [120], japonica rice [121], onion, and Nicotiana benthamiana [122].
To increase the efficiency and accuracy of Cas9, many strategies for and modifications to the enzyme have been developed. The intronized Cas9 gene has been studied and exhibited an increase in the mutation rate from 70% to 100% in comparison with the non-intronized Cas9 gene in Arabidopsis thaliana, Nicotiana benthamiana, and Cartharanthus roseus [123]. Because the ideal growth condition for Streptococcus pyrogenes is 40 °C, the efficiency of on-target mutations after incubation cycles at 37 °C was increased in Arabidopsis thaliana and citrus [124]. However, this is not an ideal temperature for the growth of many plants. The Cas enzyme, which works at lower temperatures and even responds to light, has been developed [125].
These innovative DNA delivery methods for increased efficiency can overcome the limitations of Agrobacterium-based genome editing and accelerate the process of gene editing-based plant breeding.

4.3. Broadening Gene Targets and Crops

Another challenge is found in editing the genome of polyploid species, since many studies and products obtained have focused on diploid species or model plants. In polyploid species, mutations that cause a loss of function are usually inefficient because of genetic redundancy [126,127]. In the case of CRISPR/Cas9, multiple gRNAs can be used, and mutations in different alleles can be introduced. In addition to wheat, this technique can be successfully applied to cotton [128], strawberries [129], and potatoes [130].

5. Conclusions

ZFNs, TALENs, and CRISPR/Cas systems have been used for different purposes in modern agriculture, such as obtaining plants with tolerance to abiotic stresses, such as temperature, drought, and salinity stresses. With the emergence of CRISPR/Cas, these processes have become more rapid and accurate. In recent years, many advances have been made in genome-editing technologies, such as the modification of known structures or the discovery of new enzymes or methodologies. With the advancement of biotechnology, new tools can be explored and improved to overcome barriers and the surprising progress in the genome-editing system in plants.

Author Contributions

Writing, original draft preparation, X.L., S.X., M.B.F.-A., S.Y., T.I. and M.K.; writing, review and editing, K.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Program on Open Innovation Platform with Enterprises, Research Institute, and Academia, Japan Science and Technology Agency JST-OPERA (JPMJOP1851), JST-SPRING (JPMJSP2124), the JSPS Grant-in-Aid for Scientific Research (22H02295), and a Cooperative Research Grant of the Plant Transgenic Design Initiative (PTraD) by the Tsukuba-Plant Innovation Research Center, University of Tsukuba.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.B.F.A. acknowledges the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan for providing a scholarship to conduct the current research under the Japan Leader Empowerment Program on Global Supports towards Regional Revitalization (JLEP) for a doctoral course.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Meteorological Organization. State of Climate in 2021: Extreme Events and Major Impacts; World Meteorological Organization: Geneva, Switzerland, 2021. [Google Scholar]
  2. Schmidhuber, J.; Tubiello, F.N. Global food security under climate change. Proc. Natl. Acad. Sci. USA 2007, 104, 19703–19708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Darwin, R.; Kennedy, D. Economic effects of CO2 fertilization of crops: Transforming changes in yield into changes in supply. Environ. Model. Assess. 2000, 5, 157–168. [Google Scholar] [CrossRef]
  4. Porter, J.R.; Xie, L. Food Security and Food Production Systems. In Climate Change 2014—Impacts, Adaptation and Vulnerability: Part A: Global and Sectoral Aspects: Working Group II Contribution to the IPCC Fifth Assessment Report: Volume 1: Global and Sectoral Aspects; Intergovernmental Panel on Climate Change, Ed.; Cambridge University Press: Cambridge, UK, 2014; Volume 1, pp. 485–534. [Google Scholar]
  5. Rane, J.; Singh, A.K.; Kumar, M.; Boraiah, K.M.; Meena, K.K.; Pradhan, A.; Prasad, P.V.V. The adaptation and tolerance of major cereals and legumes to important abiotic stresses. Int. J. Mol. Sci. 2021, 22, 12970. [Google Scholar] [CrossRef] [PubMed]
  6. Kumar, S. Abiotic stresses and their effects on plant growth, yield and nutritional quality of agricultural produce. Int. J. Sci. Food Agric. 2020, 4, 367–378. [Google Scholar] [CrossRef]
  7. Ashraf, M.; Foolad, M.R. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ. Exp. Bot. 2007, 59, 206–216. [Google Scholar] [CrossRef]
  8. Salava, H.; Thula, S.; Mohan, V.; Kumar, R.; Maghuly, F. Application of genome editing in tomato breeding: Mechanisms, advances, and prospects. Int. J. Mol. Sci. 2021, 22, 682. [Google Scholar] [CrossRef]
  9. Yamamoto, T.; Kashojiya, S.; Kamimura, S.; Kameyama, T.; Ariizumi, T.; Ezura, H.; Miura, K. Application and development of genome editing technologies to the Solanaceae plants. Plant Physiol. Bioch. 2018, 131, 37–46. [Google Scholar] [CrossRef]
  10. Xia, X.; Cheng, X.; Li, R.; Yao, J.; Li, Z.; Cheng, Y. Advances in application of genome editing in tomato and recent development of genome editing technology. Theor. App. Genet. 2021, 134, 2727–2747. [Google Scholar] [CrossRef]
  11. Mushtaq, M.; Ahmad Dar, A.; Skalicky, M.; Tyagi, A.; Bhagat, N.; Basu, U.; Bhat, B.A.; Zaid, A.; Ali, S.; Dar, T.-U.-H.; et al. CRISPR-based genome editing tools: Insights into technological breakthroughs and future challenges. Genes 2021, 12, 797. [Google Scholar] [CrossRef]
  12. Gaj, T.; Gersbach, C.A.; Barbas, C.F., 3rd. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [Green Version]
  13. Carroll, D. Genome engineering with zinc-finger nucleases. Genetics 2011, 188, 773–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhang, Y.; Massel, K.; Godwin, I.D.; Gao, C. Applications and potential of genome editing in crop improvement. Genome Biol. 2018, 19, 210. [Google Scholar] [CrossRef]
  15. Cui, Z.; Liu, H.; Zhang, H.; Huang, Z.; Tian, R.; Li, L.; Fan, W.; Chen, Y.; Chen, L.; Zhang, S.; et al. The comparison of ZFNs, TALENs, and SpCas9 by GUIDE-seq in HPV-targeted gene therapy. Mol. Ther. Nucleic. Acids. 2021, 26, 1466–1478. [Google Scholar] [CrossRef] [PubMed]
  16. Lloyd, A.; Plaisier, C.L.; Carroll, D.; Drews, G.N. Targeted mutagenesis using zinc-finger nucleases in Arabidopsis. Proc. Natl. Acad. Sci. USA 2005, 102, 2232–2237. [Google Scholar] [CrossRef] [Green Version]
  17. Saha, S.K.; Saikot, F.K.; Rahman, M.S.; Jamal, M.A.H.M.; Rahman, S.M.K.; Islam, S.M.R.; Kim, K.H. Programmable molecular scissors: Applications of a new tool for genome editing in biotech. Mol. Ther. Nucleic. Acids. 2019, 14, 212–238. [Google Scholar] [CrossRef] [Green Version]
  18. Makarova, K.S.; Haft, D.H.; Barrangou, R.; Brouns, S.J.J.; Charpentier, E.; Horvath, P.; Moineau, S.; Mojica, F.J.M.; Wolf, Y.I.; Yakunin, A.F.; et al. Evolution and classification of the CRISPR–Cas systems. Nat. Rev. Microbiol. 2011, 9, 467–477. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, H.; La Russa, M.; Qi, L.S. CRISPR/Cas9 in genome editing and beyond. Annu. Rev. Biochem. 2016, 85, 227–264. [Google Scholar] [CrossRef] [Green Version]
  20. Kazi, T.A.; Biswas, S.R. CRISPR/dCas system as the modulator of gene expression. Prog. Mol. Biol. Transl. Sci. 2021, 178, 99–122. [Google Scholar]
  21. Park, K.H.; An, Y.; Woo, E.J. In vitro assembly of thermostable Csm complex in CRISPR-Cas type III/A system. Methods Enzymol. 2019, 616, 173–189. [Google Scholar]
  22. Matres, J.M.; Hilscher, J.; Datta, A.; Armario-Nájera, V.; Baysal, C.; He, W.; Huang, X.; Zhu, C.; Valizadeh-Kamran, R.; Trijatmiko, K.R.; et al. Genome editing in cereal crops: An overview. Transgenic. Res. 2021, 30, 461–498. [Google Scholar] [CrossRef]
  23. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; van der Oost, J.; Regev, A.; et al. Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas system. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef] [Green Version]
  24. Osakabe, K.; Wada, N.; Miyaji, T.; Murakami, E.; Marui, K.; Ueta, R.; Hashimoto, R.; Abe-Hara, C.; Kong, B.; Yano, K.; et al. Genome editing in plants using CRISPR type I-D nuclease. Commun. Biol. 2020, 3, 648. [Google Scholar] [CrossRef] [PubMed]
  25. Ming, M.; Ren, Q.; Pan, C.; He, Y.; Zhang, Y.; Liu, S.; Zhong, Z.; Wang, J.; Malzahn, A.A.; Wu, J.; et al. CRISPR-Cas12b enables efficient plant genome engineering. Nat. Plants 2020, 6, 202–208. [Google Scholar] [CrossRef] [PubMed]
  26. Aman, R.; Ali, Z.; Butt, H.; Mahas, A.; Aljedaani, F.; Khan, M.Z.; Ding, S.; Mahfouz, M. RNA virus interference via CRISPR/Cas13a system in plants. Genome Biol. 2018, 19, 1. [Google Scholar] [CrossRef] [PubMed]
  27. Zhan, X.; Zhang, F.; Zhong, Z.; Chen, R.; Wang, Y.; Chang, L.; Bock, R.; Nie, B.; Zhang, J. Generation of virus-resistant potato plants by RNA genome targeting. Plant Biotechnol. J. 2019, 17, 1814–1822. [Google Scholar] [CrossRef] [Green Version]
  28. 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-CasΦ from huge phages is a hypercompact genome editor. Science 2020, 369, 333–337. [Google Scholar] [CrossRef]
  29. Ran, F.A.; Cong, L.; Yan, W.X.; Scott, D.A.; Gootenberg, J.S.; Kriz, A.J.; Zetsche, B.; Shalem, O.; Wu, X.; Makarova, K.S.; et al. In vivo genome editing using Staphylococcus aureus Cas9. Nature 2015, 520, 186–191. [Google Scholar] [CrossRef]
  30. Müller, 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]
  31. Kantor, A.; McClements, M.E.; MacLaren, R.E. CRISPR-Cas9 DNA base-editing and prime-editing. Int. J. Mol. Sci. 2020, 21, 6240. [Google Scholar] [CrossRef]
  32. 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]
  33. Zhao, K.; Tung, C.W.; Eizenga, G.C.; Wright, M.H.; Ali, M.L.; Price, A.H.; Norton, G.J.; Islam, M.R.; Reynolds, A.; Mezey, J.; et al. Genome-wide association mapping reveals a rich genetic architecture of complex traits in Oryza sativa. Nat. Commun. 2011, 2, 467. [Google Scholar] [CrossRef] [PubMed]
  34. Zong, Y.; Wang, Y.; Li, C.; Zhang, R.; Chen, K.; Ran, Y.; Qiu, J.-L.; Wang, D.; Gao, C. Precise base editing in rice, wheat and maize with a Cas9-cytidine deaminase fusion. Nat. Biotechnol. 2017, 35, 438–440. [Google Scholar] [CrossRef]
  35. Gao, C. Genome engineering for crop improvement and future agriculture. Cell 2021, 184, 1621–1635. [Google Scholar] [CrossRef]
  36. Armario Najera, V.; Twyman, R.M.; Christou, P.; Zhu, C. Applications of multiplex genome editing in higher plants. Curr. Opin. Chem. Biol. 2019, 59, 93–102. [Google Scholar] [CrossRef] [PubMed]
  37. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef]
  38. 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] [PubMed]
  39. 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]
  40. Nawaz, G.; Han, Y.; Usman, B.; Liu, F.; Qin, B.; Li, R. Knockout of OsPRP1, a gene encoding proline-rich protein, confers enhanced cold sensitivity in rice (Oryza sativa L.) at the seedling stage. 3 Biotech 2019, 9, 254. [Google Scholar] [CrossRef]
  41. 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] [Green Version]
  42. Zhao, C.; Zhang, Z.; Xie, S.; Si, T.; Li, Y.; Zhu, J.K. Mutational evidence for the critical role of CBF transcription factors in cold acclimation in Arabidopsis. Plant Physiol. 2016, 171, 2744–2759. [Google Scholar] [CrossRef] [Green Version]
  43. Jia, Y.; Ding, Y.; Shi, Y.; Zhang, X.; Gong, Z.; Yang, S. The cbfs triple mutants reveal the essential functions of CBFs in cold acclimation and allow the definition of CBF regulons in Arabidopsis. New. Phytol. 2016, 212, 345–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wang, L.; Zhao, R.; Zheng, Y.; Chen, L.; Li, R.; Ma, J.; Hong, X.; Ma, P.; Sheng, J.; Shen, L. SlMAPK1/2/3 and antioxidant enzymes are associated with H2O2-induced chilling tolerance in tomato plants. J. Agric. Food Chem. 2017, 65, 6812–6820. [Google Scholar] [CrossRef] [PubMed]
  45. Li, R.; Zhang, L.; Wang, L.; Chen, L.; Zhao, R.; Sheng, J.; Shen, L. Reduction of tomato-plant chilling tolerance by CRISPR–Cas9-mediated SlCBF1 mutagenesis. J. Agric. Food Chem. 2018, 66, 9042–9051. [Google Scholar] [CrossRef] [PubMed]
  46. Shen, C.; Que, Z.; Lu, Q.; Liu, T.; Li, S.; Zou, J.; Chen, G. The Rice Annexin Gene OsAnn5 Is a Positive Regulator of Cold Stress Tolerance at the Seedling Stage; Research Square: Durham, NC, USA, 2020. [Google Scholar]
  47. Shen, C.; Que, Z.; Xia, Y.; Tang, N.; Li, D.; He, R.; Cao, M. Knock out of the annexin gene OsAnn3 via CRISPR/Cas9-mediated genome editing decreased cold tolerance in rice. J. Plant Biol. 2017, 60, 539–547. [Google Scholar] [CrossRef]
  48. Qiu, Z.; Kang, S.; He, L.; Zhao, J.; Zhang, S.; Hu, J.; Zeng, D.; Zhang, G.; Dong, G.; Gao, Z.; et al. The newly identified heat-stress sensitive albino 1 gene affects chloroplast development in rice. Plant Sci. 2018, 267, 168–179. [Google Scholar] [CrossRef]
  49. Hu, Z.; Li, J.; Ding, S.; Cheng, F.; Li, X.; Jiang, Y.; Yu, J.; Foyer, C.H.; Shi, K. The protein kinase CPK28 phosphorylates ascorbate peroxidase and enhances thermotolerance in tomato. Plant Physiol. 2021, 186, 1302–1317. [Google Scholar] [CrossRef]
  50. 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]
  51. Yin, Y.; Qin, K.; Song, X.; Zhang, Q.; Zhou, Y.; Xia, X.; Yu, J. BZR1 transcription factor regulates heat stress tolerance through FERONIA receptor-like kinase-mediated reactive oxygen species signaling in tomato. Plant Cell Physiol. 2018, 59, 2239–2254. [Google Scholar] [CrossRef]
  52. Wang, B.; Zhong, Z.; Wang, X.; Han, X.; Yu, D.; Wang, C.; Song, W.; Zheng, X.; Chen, C.; Zhang, Y. Knockout of the OsNAC006 transcription factor causes drought and heat sensitivity in rice. Int. J. Mol. Sci. 2020, 21, 2288. [Google Scholar] [CrossRef] [Green Version]
  53. 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]
  54. Ueta, R.; Abe, C.; Watanabe, T.; Sugano, S.S.; Ishihara, R.; Ezura, H.; Osakabe, Y.; Osakabe, K. Rapid breeding of parthenocarpic tomato plants using CRISPR/Cas9. Sci. Rep. 2017, 7, 507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kashojiya, S.; Lu, Y.; Takayama, M.; Komatsu, H.; Minh, L.H.T.; Nishida, K.; Shirasawa, K.; Miura, K.; Nonaka, S.; Masuda, J.I.; et al. Modification of tomato breeding traits and plant hormone signaling by target-AID, the genome-editing system inducing efficient nucleotide substitution. Hortic. Res. 2022, 9, uhab004. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, L.; Zhang, J.; Xu, J.; Li, Y.; Guo, L.; Wang, Z.; Zhang, X.; Zhao, B.; Guo, Y.-D.; Zhang, N. CRISPR/Cas9 targeted mutagenesis of SlLBD40, a lateral organ boundaries domain transcription factor, enhances drought tolerance in tomato. Plant Sci. 2020, 301, 110683. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, M.; Zhu, X.; Liu, X.; Wu, C.; Yu, C.; Hu, G.; Chen, L.; Chen, R.; Bouzayen, M.; Zouine, M.; et al. Knockout of auxin response factor SlARF4 improves tomato resistance to water deficit. Int. J. Mol. Sci. 2021, 22, 3347. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, S.; Zhang, N.; Zhou, G.; Hussain, S.; Ahmed, S.; Tian, H.; Wang, S. Knockout of the entire family of AITR genes in Arabidopsis leads to enhanced drought and salinity tolerance without fitness costs. BMC Plant Biol. 2021, 21, 137. [Google Scholar] [CrossRef]
  59. Osakabe, Y.; Watanabe, T.; Sugano, S.S.; Ueta, R.; Ishihara, R.; Shinozaki, K.; Osakabe, K. Optimization of CRISPR/Cas9 genome editing to modify abiotic stress responses in plants. Sci. Rep. 2016, 6, 26685. [Google Scholar] [CrossRef] [Green Version]
  60. Roca Paixão, J.F.; Gillet, F.-X.; Ribeiro, T.P.; Bournaud, C.; Lourenço-Tessutti, I.T.; Noriega, D.D.; de Melo, B.P.; de Almeida-Engler, J.; Grossi-de-Sa, M.F. Improved drought stress tolerance in Arabidopsis by CRISPR/dCas9 fusion with a histone acetyltransferase. Sci. Rep. 2019, 9, 8080. [Google Scholar] [CrossRef] [Green Version]
  61. Shi, J.; Gao, H.; Wang, H.; Lafitte, H.R.; Archibald, R.L.; Yang, M.; Hakimi, S.M.; Mo, H.; Habben, J.E. ARGOS8 variants generated by CRISPR-Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnol. J. 2017, 15, 207–216. [Google Scholar] [CrossRef] [Green Version]
  62. 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]
  63. Liao, S.; Qin, X.; Luo, L.; Han, Y.; Wang, X.; Usman, B.; Nawaz, G.; Zhao, N.; Liu, Y.; Li, R. CRISPR/Cas9-induced mutagenesis of Semi-Rolled Leaf1,2 confers curled leaf phenotype and drought tolerance by influencing protein expression patterns and ROS scavenging in rice (Oryza sativa L.). Agronomy 2019, 9, 728. [Google Scholar] [CrossRef] [Green Version]
  64. Li, R.; Liu, C.; Zhao, R.; Wang, L.; Chen, L.; Yu, W.; Zhang, S.; Sheng, J.; Shen, L. CRISPR/Cas9-mediated SlNPR1 mutagenesis reduces tomato plant drought tolerance. BMC Plant Biol. 2019, 19, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. 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] [PubMed]
  66. Du, Y.T.; Zhao, M.J.; Wang, C.T.; Gao, Y.; Wang, Y.X.; Liu, Y.W.; Chen, M.; Chen, J.; Zhou, Y.B.; Xu, Z.S.; et al. Identification and characterization of GmMYB118 responses to drought and salt stress. BMC Plant Biol. 2018, 18, 320. [Google Scholar] [CrossRef] [PubMed]
  67. Qin, Q.; Wang, Y.; Huang, L.; Du, F.; Zhao, X.; Li, Z.; Wang, W.; Fu, B. A U-box E3 ubiquitin ligase OsPUB67 is positively involved in drought tolerance in rice. Plant Mol. Biol. 2020, 102, 89–107. [Google Scholar] [CrossRef]
  68. Lou, D.; Wang, H.; Liang, G.; Yu, D. OsSAPK2 confers abscisic acid sensitivity and tolerance to drought stress in rice. Front. Plant Sci. 2017, 8, 993. [Google Scholar] [CrossRef] [Green Version]
  69. Zhang, A.; Liu, Y.; Wang, F.; Li, T.; Chen, Z.; Kong, D.; Bi, J.; Zhang, F.; Luo, X.; Wang, J.; et al. Enhanced rice salinity tolerance via CRISPR/Cas9-targeted mutagenesis of the OsRR22 gene. Mol. Breed. 2019, 39, 47. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, X.; Ren, P.; Ji, L.; Zhu, B.; Xie, G. OsVDE, a xanthophyll cycle key enzyme, mediates abscisic acid biosynthesis and negatively regulates salinity tolerance in rice. Planta 2021, 255, 6. [Google Scholar] [CrossRef]
  71. Santosh Kumar, V.V.; Verma, R.K.; Yadav, S.K.; Yadav, P.; Watts, A.; Rao, M.V.; Chinnusamy, V. CRISPR-Cas9 mediated genome editing of drought and salt tolerance (OsDST) gene in indica mega rice cultivar MTU1010. Physiol. Mol. Biol. Plants 2020, 26, 1099–1110. [Google Scholar] [CrossRef]
  72. 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]
  73. 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]
  74. Bouzroud, S.; Gasparini, K.; Hu, G.; Barbosa, M.A.M.; Rosa, B.L.; Fahr, M.; Bendaou, N.; Bouzayen, M.; Zsögön, A.; Smouni, 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]
  75. Kim, S.-T.; Choi, M.; Bae, S.-J.; Kim, J.-S. The functional association of ACQOS/VICTR with salt stress resistance in Arabidopsis thaliana was confirmed by CRISPR-Mediated mutagenesis. Int. J. Mol. Sci. 2021, 22, 11389. [Google Scholar] [CrossRef] [PubMed]
  76. Zheng, M.; Lin, J.; Liu, X.; Chu, W.; Li, J.; Gao, Y.; An, K.; Song, W.; Xin, M.; Yao, Y.; et al. Histone acetyltransferase TaHAG1 acts as a crucial regulator to strengthen salt tolerance of hexaploid wheat. Plant Physiol. 2021, 186, 1951–1969. [Google Scholar] [CrossRef] [PubMed]
  77. Makhotenko, A.V.; Khromov, A.V.; Snigir, E.A.; Makarova, S.S.; Makarov, V.V.; Suprunova, T.P.; Kalinina, N.O.; Taliansky, M.E. Functional analysis of coilin in virus resistance and stress tolerance of potato Solanum tuberosum using CRISPR-Cas9 editing. Dokl. Biochem. Biophys. 2019, 484, 88–91. [Google Scholar] [CrossRef] [PubMed]
  78. 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] [PubMed]
  79. Liu, L.; Ji, H.; An, J.; Shi, K.; Ma, J.; Liu, B.; Tang, L.; Cao, W.; Zhu, Y. Response of biomass accumulation in wheat to low-temperature stress at jointing and booting stages. Environ. Exp. Bot. 2019, 157, 46–57. [Google Scholar] [CrossRef]
  80. Zhang, X.; Fowler, S.G.; Cheng, H.; Lou, Y.; Rhee, S.Y.; Stockinger, E.J.; Thomashow, M.F. Freezing-sensitive tomato has a functional CBF cold response pathway, but a CBF regulon that differs from that of freezing-tolerant Arabidopsis. Plant J. 2004, 39, 905–919. [Google Scholar] [CrossRef] [PubMed]
  81. Hussain, M.; Farooq, S.; Hasan, W.; Ul-Allah, S.; Tanveer, M.; Farooq, M.; Nawaz, A. Drought stress in sunflower: Physiological effects and its management through breeding and agronomic alternatives. Agric. Water Manag. 2018, 201, 152–166. [Google Scholar] [CrossRef]
  82. Nadeem, M.; Li, J.; Yahya, M.; Wang, M.; Ali, A.; Cheng, A.; Wang, X.; Ma, C. Grain legumes and fear of salt stress: Focus on mechanisms and management strategies. Int. J. Mol. Sci. 2019, 20, 799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kagale, S.; Divi, U.K.; Krochko, J.E.; Keller, W.A.; Krishna, P. Brassinosteroid confers tolerance in Arabidopsis thaliana and Brassica napus to a range of abiotic stresses. Planta 2007, 225, 353–364. [Google Scholar] [CrossRef]
  84. Jung, J.H.; Domijan, M.; Klose, C.; Biswas, S.; Ezer, D.; Gao, M.; Khattak, A.K.; Box, M.S.; Charoensawan, V.; Cortijo, S.; et al. Phytochromes function as thermosensors in Arabidopsis. Science 2016, 354, 886–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Legris, M.; Klose, C.; Burgie, E.S.; Rojas, C.C.R.; Neme, M.; Hiltbrunner, A.; Wigge, P.A.; Schäfer, E.; Vierstra, R.D.; Casal, J.J. Phytochrome B integrates light and temperature signals in Arabidopsis. Science 2016, 354, 897–900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Arico, D.; Legris, M.; Castro, L.; Garcia, C.F.; Laino, A.; Casal, J.J.; Mazzella, M.A. Neighbour signals perceived by phytochrome B increase thermotolerance in Arabidopsis. Plant Cell Environ. 2019, 42, 2554–2566. [Google Scholar] [CrossRef] [PubMed]
  87. Abdellatif, I.M.Y.; Yuan, S.; Na, R.; Yoshihara, S.; Hamada, H.; Suzaki, T.; Ezura, H.; Miura, K. Functional characterization of tomato phytochrome A and B1B2 mutants in response to heat stress. Int. J. Mol. Sci. 2022, 23, 1681. [Google Scholar] [CrossRef]
  88. Wahid, A.; Gelani, S.; Ashraf, M.; Foolad, M.R. Heat tolerance in plants: An overview. Environ. Exp. Bot. 2007, 61, 199–223. [Google Scholar] [CrossRef]
  89. Alsamir, M.; Mahmood, T.; Trethowan, R.; Ahmad, N. An overview of heat stress in tomato (Solanum lycopersicum L.). Saudi J. Biol. Sci. 2021, 28, 1654–1663. [Google Scholar] [CrossRef]
  90. Karapanos, I.C.; Akoumianakis, K.A.; Olympios, C.M.; Passam, H.C. The effect of substrate, ADP and uncoupler on the respiration of tomato pollen during incubation in vitro at moderately high temperature. Sex. Plant Reprod. 2009, 22, 133–140. [Google Scholar] [CrossRef]
  91. Ruan, Y.L.; Patrick, J.W.; Bouzayen, M.; Osorio, S.; Fernie, A.R. Molecular regulation of seed and fruit set. Trends. Plant Sci. 2012, 17, 656–665. [Google Scholar] [CrossRef] [Green Version]
  92. Ingrosso, I.; Bonsegna, S.; De Domenico, S.; Laddomada, B.; Blando, F.; Santino, A.; Giovinazzo, G. Over-expression of a grape stilbene synthase gene in tomato induces parthenocarpy and causes abnormal pollen development. Plant Physiol. Biochem. 2011, 49, 1092–1099. [Google Scholar] [CrossRef]
  93. Livne, S.; Lor, V.S.; Nir, I.; Eliaz, N.; Aharoni, A.; Olszewski, N.E.; Eshed, Y.; Weiss, D. Uncovering DELLA-independent gibberellin responses by characterizing new tomato procera mutants. Plant Cell 2015, 27, 1579–1594. [Google Scholar] [CrossRef] [Green Version]
  94. Merlot, S.; Leonhardt, N.; Fenzi, F.; Valon, C.; Costa, M.; Piette, L.; Vavasseur, A.; Genty, B.; Boivin, K.; Müller, A.; et al. Constitutive activation of a plasma membrane H+-ATPase prevents abscisic acid-mediated stomatal closure. EMBO J. 2007, 26, 3216–3226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zörb, C.; Geilfus, C.-M.; Dietz, K.-J. Salinity and crop yield. Plant Biol. 2019, 21, 31–38. [Google Scholar] [CrossRef] [PubMed]
  96. Bailey-Serres, J.; Parker, J.E.; Ainsworth, E.A.; Oldroyd, G.E.D.; Schroeder, J.I. Genetic strategies for improving crop yields. Nature 2019, 575, 109–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Biswas, S.; Zhang, D.; Shi, J. CRISPR/Cas systems: Opportunities and challenges for crop breeding. Plant Cell Rep. 2021, 40, 979–998. [Google Scholar] [CrossRef] [PubMed]
  98. Sedeek, K.E.M.; Mahas, A.; Mahfouz, M. Plant genome engineering for targeted improvement of crop traits. Front. Plant Sci. 2019, 10, 114. [Google Scholar] [CrossRef] [Green Version]
  99. Qiu, X.Y.; Zhu, L.Y.; Zhu, C.S.; Ma, J.X.; Hou, T.; Wu, X.M.; Xie, S.S.; Min, L.; Tan, D.A.; Zhang, D.Y.; et al. Highly effective and low-cost microRNA detection with CRISPR-Cas9. ACS Synth. Biol. 2018, 7, 807–813. [Google Scholar] [CrossRef]
  100. Swarts, D.C.; Jinek, M. Cas9 versus Cas12a/Cpf1: Structure-function comparisons and implications for genome editing. Wiley Interdiscip. Rev. RNA 2018, 9, e1481. [Google Scholar] [CrossRef]
  101. 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]
  102. Lassoued, R.; Phillips, P.W.B.; Macall, D.M.; Hesseln, H.; Smyth, S.J. Expert opinions on the regulation of plant genome editing. Plant Biotechnol. J. 2021, 19, 1104–1109. [Google Scholar] [CrossRef] [PubMed]
  103. Nagamine, A.; Ezura, H. Genome editing for improving crop nutrition. Front. Genome Ed. 2022, 4, 850104. [Google Scholar] [CrossRef]
  104. Yang, B. Grand challenges in genome editing in plants. Front. Genome Ed. 2020, 2, 2. [Google Scholar] [CrossRef] [PubMed]
  105. Ruffell, D. The EU Court of Justice extends the GMO Directive to gene-edited organisms. FEBS Lett. 2018, 592, 3653–3657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Shew, A.M.; Nalley, L.L.; Snell, H.A.; Nayga, R.M.; Dixon, B.L. CRISPR versus GMOs: Public acceptance and valuation. Glob. Food Sec. 2018, 19, 71–80. [Google Scholar] [CrossRef]
  107. Liu, H.; Zhang, B. Virus-based CRISPR/Cas9 genome editing in plants. Trends Genet. 2020, 36, 810–813. [Google Scholar] [CrossRef]
  108. Ma, X.; Zhang, X.; Liu, H.; Li, Z. Highly efficient DNA-free plant genome editing using virally delivered CRISPR–Cas9. Nat. Plants 2020, 6, 773–779. [Google Scholar] [CrossRef] [PubMed]
  109. Ariga, H.; Toki, S.; Ishibashi, K. Potato virus X vector-mediated DNA-free genome editing in plants. Plant Cell Physiol. 2020, 61, 1946–1953. [Google Scholar] [CrossRef]
  110. Sharma, P.; Lew, T.T.S. Principles of nanoparticle design for genome editing in plants. Front. Genome Ed. 2022, 4, 846624. [Google Scholar] [CrossRef]
  111. Mahmoud, L.M.; Kaur, P.; Stanton, D.; Grosser, J.W.; Dutt, M. A cationic lipid mediated CRISPR/Cas9 technique for the production of stable genome edited citrus plants. Plant Methods 2022, 18, 33. [Google Scholar] [CrossRef]
  112. Maher, M.F.; Nasti, R.A.; Vollbrecht, M.; Starker, C.G.; Clark, M.D.; Voytas, D.F. Plant gene editing through de novo induction of meristems. Nat. Biotechnol. 2020, 38, 84–89. [Google Scholar] [CrossRef]
  113. He, Y.; Zhao, Y. Technological breakthroughs in generating transgene-free and genetically stable CRISPR-edited plants. aBIOTECH 2020, 1, 88–96. [Google Scholar] [CrossRef] [Green Version]
  114. Yuan, S.; Kawasaki, S.; Abdellatif, I.M.Y.; Nishida, K.; Kondo, A.; Ariizumi, T.; Ezura, H.; Miura, K. Efficient base editing in tomato using a highly expressed transient system. Plant Cell Rep. 2021, 40, 667–676. [Google Scholar] [CrossRef]
  115. Veillet, F.; Perrot, L.; Chauvin, L.; Kermarrec, M.P.; Guyon-Debast, A.; Chauvin, J.E.; Nogué, F.; Mazier, M. Transgene-free genome editing in tomato and potato plants using Agrobacterium-mediated delivery of a CRISPR/Cas9 cytidine base editor. Int. J. Mol. Sci. 2019, 20, 402. [Google Scholar] [CrossRef] [Green Version]
  116. Bánfalvi, Z.; Csákvári, E.; Villányi, V.; Kondrák, M. Generation of transgene-free PDS mutants in potato by Agrobacterium-mediated transformation. BMC Biotechnol. 2020, 20, 25. [Google Scholar] [CrossRef]
  117. Chen, L.; Li, W.; Katin-Grazzini, L.; Ding, J.; Gu, X.; Li, Y.; Gu, T.; Wang, R.; Lin, X.; Deng, Z.; et al. A method for the production and expedient screening of CRISPR/Cas9-mediated non-transgenic mutant plants. Hortic. Res. 2018, 5, 13. [Google Scholar] [CrossRef]
  118. Imai, R.; Hamada, H.; Liu, Y.; Linghu, Q.; Kumagai, Y.; Nagira, Y.; Miki, R.; Taoka, N. In planta particle bombardment (iPB): A new method for plant transformation and genome editing. Plant Biotechnol. 2020, 37, 171–176. [Google Scholar] [CrossRef]
  119. Hamada, H.; Liu, Y.; Nagira, Y.; Miki, R.; Taoka, N.; Imai, R. Biolistic-delivery-based transient CRISPR/Cas9 expression enables in planta genome editing in wheat. Sci. Rep. 2018, 8, 14422. [Google Scholar] [CrossRef]
  120. Liu, G.; Li, J.; Godwin, I.D. Genome editing by CRISPR/Cas9 in sorghum through biolistic bombardment. Methods. Mol. Biol. 2019, 1931, 169–183. [Google Scholar]
  121. Molina-Risco, M.; Ibarra, O.; Faion-Molina, M.; Kim, B.; Septiningsih, E.M.; Thomson, M.J. Optimizing Agrobacterium-mediated transformation and CRISPR-Cas9 gene editing in the tropical japonica rice variety presidio. Int. J. Mol. Sci. 2021, 22, 10909. [Google Scholar] [CrossRef]
  122. Miller, K.; Eggenberger, A.L.; Lee, K.; Liu, F.; Kang, M.; Drent, M.; Ruba, A.; Kirscht, T.; Wang, K.; Jiang, S. An improved biolistic delivery and analysis method for evaluation of DNA and CRISPR-Cas delivery efficacy in plant tissue. Sci. Rep. 2021, 11, 7695. [Google Scholar] [CrossRef]
  123. Grützner, R.; Martin, P.; Horn, C.; Mortensen, S.; Cram, E.J.; Lee-Parsons, C.W.T.; Stuttmann, J.; Marillonnet, S. High-efficiency genome editing in plants mediated by a Cas9 gene containing multiple introns. Plant Commun. 2021, 2, 100135. [Google Scholar] [CrossRef]
  124. LeBlanc, C.; Zhang, F.; Mendez, J.; Lozano, Y.; Chatpar, K.; Irish, V.F.; Jacob, Y. Increased efficiency of targeted mutagenesis by CRISPR/Cas9 in plants using heat stress. Plant J. 2018, 93, 377–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Richter, F.; Fonfara, I.; Bouazza, B.; Schumacher, C.H.; Bratovič, M.; Charpentier, E.; Möglich, A. Engineering of temperature- and light-switchable Cas9 variants. Nucleic. Acids Res. 2016, 44, 10003–10014. [Google Scholar] [CrossRef] [Green Version]
  126. 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]
  127. Arndell, T.; Sharma, N.; Langridge, P.; Baumann, U.; Watson-Haigh, N.S.; Whitford, R. gRNA validation for wheat genome editing with the CRISPR-Cas9 system. BMC Biotechnol. 2019, 19, 71. [Google Scholar] [CrossRef] [PubMed]
  128. Gao, W.; Long, L.; Tian, X.; Xu, F.; Liu, J.; Singh, P.K.; Botella, J.R.; Song, C. Genome editing in cotton with the CRISPR/Cas9 system. Front. Plant Sci. 2017, 8, 1364. [Google Scholar] [CrossRef] [Green Version]
  129. Zhou, J.; Wang, G.; Liu, Z. Efficient genome editing of wild strawberry genes, vector development and validation. Plant Biotechnol. J. 2018, 16, 1868–1877. [Google Scholar] [CrossRef] [Green Version]
  130. Johansen, I.E.; Liu, Y.; Jørgensen, B.; Bennett, E.P.; Andreasson, E.; Nielsen, K.L.; Blennow, A.; Petersen, B.L. High efficacy full allelic CRISPR/Cas9 gene editing in tetraploid potato. Sci. Rep. 2019, 9, 17715. [Google Scholar] [CrossRef]
Figure 1. (A) A scheme of the TALEN architecture and genome-editing mechanism. Transcription activator-like effector nucleases (TALEN) are chimeric protein that works in pairs composed by a TAL effector DNA-binding domain merged to a nuclease domain from the FokI restriction enzyme. The TALEs have highly conserved and repetitive peptide modules containing up to 34 amino acids (represented by the colored bars). Each TALE repeat specifically recognizes one of the nucleotide bases, and multiple TALE repeats are combined to target a specific DNA sequence and generate a DNA double-strand break by FokI action within the intervening spacer region. (B) Overview of CRISPR/Cas9-mediated genome-editing mechanism for non-homologous end joining (NHEJ) repair. Guide RNA (gRNA) is designed to recognize the target sequence located upstream of the protospacer-associated motif (PAM), NGG in the case of CRISPR/Cas9, which serves as a binding signal for Cas9. When nucleotide base pairing occurs (due to the annealing of the target sequence with the protospacer region of the gRNA, represented in the figure by an orange line), the Cas9 enzyme is activated, causing DNA double-strand break. The breaks activate the intracellular repair systems of the cell, which convert breaks into insertion or deletion mutations. The mutations generally cause sequence failures and generate non-functional proteins. CRISPR-associated protein 9 (Cas9); guide RNA (gRNA); protospacer adjacent motif (PAM).
Figure 1. (A) A scheme of the TALEN architecture and genome-editing mechanism. Transcription activator-like effector nucleases (TALEN) are chimeric protein that works in pairs composed by a TAL effector DNA-binding domain merged to a nuclease domain from the FokI restriction enzyme. The TALEs have highly conserved and repetitive peptide modules containing up to 34 amino acids (represented by the colored bars). Each TALE repeat specifically recognizes one of the nucleotide bases, and multiple TALE repeats are combined to target a specific DNA sequence and generate a DNA double-strand break by FokI action within the intervening spacer region. (B) Overview of CRISPR/Cas9-mediated genome-editing mechanism for non-homologous end joining (NHEJ) repair. Guide RNA (gRNA) is designed to recognize the target sequence located upstream of the protospacer-associated motif (PAM), NGG in the case of CRISPR/Cas9, which serves as a binding signal for Cas9. When nucleotide base pairing occurs (due to the annealing of the target sequence with the protospacer region of the gRNA, represented in the figure by an orange line), the Cas9 enzyme is activated, causing DNA double-strand break. The breaks activate the intracellular repair systems of the cell, which convert breaks into insertion or deletion mutations. The mutations generally cause sequence failures and generate non-functional proteins. CRISPR-associated protein 9 (Cas9); guide RNA (gRNA); protospacer adjacent motif (PAM).
Cimb 44 00182 g001
Table 1. Application of CRISPR/Cas9-mediated genome editing to elucidate genes involved in response to temperature, drought, and salinity stresses.
Table 1. Application of CRISPR/Cas9-mediated genome editing to elucidate genes involved in response to temperature, drought, and salinity stresses.
Plant SpeciesTarget GenesGene FunctionPhenotypeMode of ApplicationRef.
RiceOsPRP1Proline-rich proteinCold sensitiveMutants exhibited sensitive phenotype after treatment at 6 °C for 3 days.[40]
RiceOsMYB30Transcription factorCold tolerance, increased panicle length, enlarged grain sizeMutants exhibited tolerance phenotype after treatment at 4 °C for 5–10 days.[41]
ArabidopsisCBFsTranscription factor for abiotic stress responsesExtremely sensitive to freezingMutants exhibited sensitive phenotype after treatment at 4 °C and freezing for 7 days and freezing-sensitive phenotype after treatment at −7 °C for 1 h.[42]
Mutants exhibited sensitive phenotype after treatment at 4 °C for 50 days and freezing sensitive after treatment at −9 °C for 1 h and −10 °C for 1 h.[43]
TomatoSlCBF1Transcription factor for abiotic stress responsesMore severe chilling injury symptomsMutants exhibited sensitive phenotype after treatment at 4 °C for 7 days.[44,45]
RiceOsAnn5AnnexinCold toleranceMutants exhibited tolerance phenotype after treatment at 4–6 °C for 3 days.[46]
RiceOsAnn3AnnexinCold toleranceMutants exhibited tolerance phenotype after treatment at 4–6 °C for 3 days.[47]
RiceOsHSA1Fructokinase-like protein 2Heat sensitiveMutants exhibited tolerance phenotype after treatment at 32 °C for 60 days.[48]
TomatoSlCPK28Protein kinase, Ca2+ sensingHeat sensitive, accumulation of ROSMutants exhibited sensitive phenotype and higher H2O2 content after treatment at 45 °C for 12 h.[49]
TomatoSlMAPK3MAP kinase upregulating HSPs’/HSFs’ genes’ expressionHeat tolerance, reduction of ROS accumulationMutants exhibited tolerance phenotype and lower H2O2 and O2•− contents after treatment at 42 °C for 1 day.[50]
TomatoSlBZR1Transcription factor for brassinosteroid responseHeat toleranceMutants exhibited tolerance phenotype after treatment at 42 °C/38 °C (day/night) for 1 day.[51]
RiceOsNAC006NAC transcription factorHeat sensitiveMutants exhibited sensitive phenotype after treatment at 42 °C for 4 days.[52]
TomatoSlAGL6MADS-boxParthenocarpy, tomato fruit under heat stressMutants exhibited facultative parthenocarpy phenotype after treatment under natural heat stress for 67 days.[53]
TomatoSlIAA9Transcriptional regulator for auxin responseParthenocarpyMutants exhibited parthenocarpy phenotype.[54]
TomatoSlIAA9Transcriptional regulator for auxin responseParthenocarpyMutants exhibited parthenocarpy phenotype.[55]
TomatoSlLBD40Plant-specific transcription factorsEnhanced drought tolerance and reduced stomatal conductanceMutants showed drought-tolerant phenotype under the 10-day watering cessation treatment.[56]
TomatoSlARF4Auxin response factorsEnhanced drought tolerance and stem thicknessMutants showed drought-tolerant phenotype under the 12-day watering-off treatment.[57]
ArabidopsisAtAITR familyABA-induced transcription repressorsEnhanced drought and salt tolerance, reduced ABA sensitivityMutants showed drought-tolerant phenotype after 12-day watering off treatment and 2 days of rewatering.[58]
ArabidopsisAtOST2Stomatal opening regulatorEnhanced drought tolerance and stomatal closureMutants showed a lower water loss rate than the wild type after 5 h of normal treatment.[59]
ArabidopsisAREB1ABA-responsive element-binding proteinEnhanced drought tolerance and chlorophyll contentMutants showed drought-tolerant phenotype under 20% humidity treatment or 20-day cessation of watering.[60]
MaizeARGOS8Negative regulator of ethylene responsesEnhanced drought tolerance, increased grain yieldMutants sown on soil with only normal 15% moisture showed drought-tolerant phenotype.[61]
RiceOsERA1ABA signaling and the dehydration responseEnhanced response to drought stress through stomatal regulationMutants showed drought-tolerant phenotype under the 8-day watering-off treatment.[62]
RiceOsSRL1,2Regulation of leaf rollingEnhanced drought tolerance and ABA levelMutants showed drought-tolerant phenotype under 30-day water-deficient treatment.[63]
TomatoSlNPR1A special receptor of salicylic acid Reduced drought tolerance, increased stomatal apertureMutants showed drought-sensitive phenotype without watering for 6 consecutive days.[64]
TomatoSlMPK3Mitogen-activated protein kinasesReduced drought tolerance, severe wilting symptomMutants showed drought-sensitive phenotype without watering for 5 consecutive days.[65]
SoybeanGmMYB118MYB transcription factor familyReduced drought and salinity toleranceMutants showed drought-sensitive phenotype after 14-day no water treatment.[66]
RiceOsPUB67U-box E3 ubiquitin ligaseReduced drought toleranceMutants showed drought-sensitive phenotype after 10-day no water treatment at tillering stage.[67]
RiceOsSAPK2Osmotic stress/ABA–activated protein kinaseReduced drought tolerance, ROS scavenging was inhibitedMutants showed drought-sensitive phenotype after 7-day no water treatment.[68]
RiceOsRR22Involved in both cytokinin signal transduction and metabolismEnhanced salinity toleranceMutants showed salinity-tolerant phenotype under concentrations of 0.75% NaCl solution treatment.[69]
RiceOsVDEKey enzyme of xanthophyll cycleEnhanced salinity tolerance, reduced water lossMutants showed salinity-tolerant phenotype at 100 mM NaCl application.[70]
RiceOsDSTDrought and salt tolerance geneEnhanced salinity tolerance, showed significantly broader leaf width and enhanced leaf areaMutants showed salinity-tolerant phenotype at 200 mM NaCl application.[71]
RiceOsNAC041NAC transcription factorReduced salinity tolerance, enhanced MDA content Mutants showed salinity-sensitive phenotype at 150 mM NaCl application.[72]
TomatoSlHyPRP1A subgroup of putative plant cell wall glycoproteinsEnhanced salinity tolerance and stem lengthMutants showed salinity-tolerant phenotype at 100 mM and150 mM NaCl application.[73]
TomatoSlARF4Auxin response factorEnhanced salinity tolerance, delayed flowering, increased height and leaf curlingMutants showed salinity-tolerant phenotype at 250 mM NaCl application.[74]
ArabidopsisAtAITRABA-induced transcription repressorsEnhanced salinity tolerance, reduced ABA sensitivityMutants showed salinity-tolerant phenotype at 150 mM NaCl application.[58]
ArabidopsisACQOSA toll-interleukin1 receptor-nucleotide-binding leucine-rich repeat class proteinEnhanced salinity tolerance and chlorophyll contentMutants showed salinity-tolerant phenotype at 250 mM NaCl application.[75]
wheatTaHAG1Histone acetyltransferaseReduced salinity tolerance, more chlorotic leaves and higher Na+ content in the mutantsMutants showed salinity-sensitive phenotype at 200 mM NaCl application.[76]
PotatoCoilinA main structural protein controlling the formation, composition, and activity of subnuclear Cajal bodiesEnhanced salinity tolerance, slower yellowing and leaf fallMutants showed salinity-tolerant phenotype at 300 mM NaCl application.[77]
SoybeanGmAITRABA-induced transcription repressorsEnhanced salinity tolerance, more sensitivity to ABAMutants showed salinity-tolerant phenotype at 200 mM NaCl application.[78]
Table 2. Comparison of the delivery of the CRISPR/Cas9 expression cassette.
Table 2. Comparison of the delivery of the CRISPR/Cas9 expression cassette.
Delivery MethodCharacteristicsLimitations
Agrobacterium-based tissue culture methodAgrobacterium infects plant cells and delivers DNA, conventionalTime-consuming and laborious, regeneration protocols required
Viral delivery methodUse of virus-based vectors for transient expressionLimited loading capacity, species-specific restriction
Nanoparticle delivery methodUse of nanoparticle–DNA complex for delivery of DNADifficult to deliver the complete CRISPR/Cas9 expression cassette through the cell wall
In planta particle bombardment methodDNA-, RNA-, and/or protein-coated particles bombarding plant tissueRegeneration required
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Xu, S.; Fuhrmann-Aoyagi, M.B.; Yuan, S.; Iwama, T.; Kobayashi, M.; Miura, K. CRISPR/Cas9 Technique for Temperature, Drought, and Salinity Stress Responses. Curr. Issues Mol. Biol. 2022, 44, 2664-2682. https://doi.org/10.3390/cimb44060182

AMA Style

Li X, Xu S, Fuhrmann-Aoyagi MB, Yuan S, Iwama T, Kobayashi M, Miura K. CRISPR/Cas9 Technique for Temperature, Drought, and Salinity Stress Responses. Current Issues in Molecular Biology. 2022; 44(6):2664-2682. https://doi.org/10.3390/cimb44060182

Chicago/Turabian Style

Li, Xiaohan, Siyan Xu, Martina Bianca Fuhrmann-Aoyagi, Shaoze Yuan, Takeru Iwama, Misaki Kobayashi, and Kenji Miura. 2022. "CRISPR/Cas9 Technique for Temperature, Drought, and Salinity Stress Responses" Current Issues in Molecular Biology 44, no. 6: 2664-2682. https://doi.org/10.3390/cimb44060182

Article Metrics

Back to TopTop