Next Article in Journal
Transcriptional Profiles of Mitochondrial Dynamics Markers Are Disturbed in Adrenal Glands of Stressed Adult Male Rats
Next Article in Special Issue
Plant Biotic and Abiotic Stresses
Previous Article in Journal
New Biomarker Combination Related to Oxidative Stress and Inflammation in Primary Open-Angle Glaucoma
Previous Article in Special Issue
Screening and Selection of Drought-Tolerant High-Yielding Chickpea Genotypes Based on Physio-Biochemical Selection Indices and Yield Trials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genome Editing and Improvement of Abiotic Stress Tolerance in Crop Plants

1
Department of Genetics & Plant Breeding, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
2
Department of Plant Molecular Biology & Biotechnology, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
3
Directorate of Research Services, Jawaharlal Nehru Krishi Vishwa Vidyalaya, Jabalpur 482004, India
4
Regional Agricultural Research Station, Sagar 470001, India
*
Authors to whom correspondence should be addressed.
Life 2023, 13(7), 1456; https://doi.org/10.3390/life13071456
Submission received: 30 May 2023 / Revised: 25 June 2023 / Accepted: 26 June 2023 / Published: 27 June 2023
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Genome editing aims to revolutionise plant breeding and could assist in safeguarding the global food supply. The inclusion of a 12–40 bp recognition site makes mega nucleases the first tools utilized for genome editing and first generation gene-editing tools. Zinc finger nucleases (ZFNs) are the second gene-editing technique, and because they create double-stranded breaks, they are more dependable and effective. ZFNs were the original designed nuclease-based approach of genome editing. The Cys2-His2 zinc finger domain’s discovery made this technique possible. Clustered regularly interspaced short palindromic repeats (CRISPR) are utilized to improve genetics, boost biomass production, increase nutrient usage efficiency, and develop disease resistance. Plant genomes can be effectively modified using genome-editing technologies to enhance characteristics without introducing foreign DNA into the genome. Next-generation plant breeding will soon be defined by these exact breeding methods. There is abroad promise that genome-edited crops will be essential in the years to come for improving the sustainability and climate-change resilience of food systems. This method also has great potential for enhancing crops’ resistance to various abiotic stressors. In this review paper, we summarize the most recent findings about the mechanism of abiotic stress response in crop plants and the use of the CRISPR/Cas mediated gene-editing systems to improve tolerance to stresses including drought, salinity, cold, heat, and heavy metals.

1. Introduction

By the end of the year 2050, the world population is anticipated to reach up to 10 billion [1]. In this situation, increasing food crop production by 60% over the coming decades is necessary to ensure global food security [1,2]. To sustainably increased food production, additional integration of all developed relevant techniques, such as genomics, genome editing (GE), artificial intelligence, and deep learning, will be necessary [3,4]. Crop modification methods have a long history and have been used ever since the first agricultural plants were domesticated. Since then, other new methods have been created and are being developed to boost crop production and economic value even more. Traditional crop breeding techniques in the 20th century either relied on naturally occurring mutations or on mutagenesis that was created artificially [5]. Genetic research has traditionally focused on the identification and assessment of spontaneous mutations. Scientists were reliant on each other and showed that radiation or chemical treatment could increase the rate of mutagenesis [6,7]. Later approaches, suchas radiation and chemical mutagenesis, altered the genome at random sites by inserting transposon motifs that may be induced in some animals. However, a fundamental disadvantage of conventional breeding methods is the length of time needed to breed new varieties of any crops with the required agronomic characteristics. The duration of the growing season and the maturity level of the plants (particularly long-period growers, such as trees), as well as various stages of crossing, selection, and testing during the breeding process, all have an impact on this [8]. The plant genome cannot be targeted using conventional techniques for chemical and physical mutagenesis or natural mutations. Using genetic engineering, better plants and animals may be developed more quickly [5].
The first genetically modified (GM) crops were released for sale in 1996 [9]. Generations of GM crops up to now have relied on the genome’s random insertion of new DNA sequences. The possibility that the inserted gene may affect or impede the activity of other crucial nearby genes has been raised as a concern regarding this approach. In addition, public anxiety regarding GM crops is increased when talking about the introduction of ‘alien’ genes from distantly related organisms, which is thought to be ‘unnatural’ despite mounting evidence to the contrary [10,11].
The creation and use of DNA-based markers at the turn of the twenty-first century has made it possible to reduce significantly the time needed to generate new lines and varieties of agricultural crops [10,11,12,13]. All these factors have greatly helped the development of focused GE methods [14,15,16,17]. In yeast and mice, the first targeted genetic alterations were created in the 1970s and 1980s [6,8]. This gene targeting was based on the homologous recombination process, which was extremely accurate.
RNA interference (RNAi) was one of the first GE technologies [5,18,19]. Even though this technology has been successfully used in functional genomics and plant breeding [20,21,22], it has several drawbacks, including the unlimited insertion site of an RNAi construction into the genome and partial gene function suppression [5].
This is a marvelous time for genetics, due to advances in genetic analysis and genetic manipulation. Genome editing, the most recent crop-enhancement method, allows precise changes of the plant genome by deleting undesired genes or enabling genes to acquire new functions [23]. Numerous crops’ genomes have been sequenced, and improvements in genome-editing techniques have made it possible to breed for desired features. To sustainably increase food production, additional integration of all developed relevant techniques, such as genomics, genome editing (GE), artificial intelligence, and deep learning, is necessary [24].
Advanced biotechnological methods are made possible by genome-editing tools, allowing for precise and effective targeted modification of an organism’s genome. Several novel tools for genome or gene editing are available to enable researchers to modify genomic sequences precisely [25]. These techniques facilitate novel insights into the functional genomics of an organism and enable us to alter the regulation of gene expression patterns in a pre-determined region. Because of accurate DNA manipulation, genome-editing technologies, for instance, CRISPR/Cas9 (clustered regularly interspaced short palindromic repeats/CRISPR-associated systems), TALENs (transcription activator-like effector nucleases), CRISPR/Cas12a (Cpf1, CRISPR from Prevotella and Francisella1), and Cas9-derived DNA base editors, provide unprecedented advancements in genome engineering. As a result, this technology is a powerful tool that can be employed to secure the global food supply [26].
Genome editing was first proposed by Capecchi [27] in the 1980s. This method allows for the removal, modification, or addition of genetic material at specified genomic locations. Even though current GE technologies are substantially more accurate than traditional mutagenesis [28,29], the biggest barrier here is still the legitimacy of GE crops. Assessing the biosafety of such crops is a unique difficulty because it is impossible to predict the effects of single base alterations following the application of ODM and BEs [30,31].
The primary elements that affect plant growth and reduce agricultural productivity are biotic stressors [32,33] such as disease and insect pests, along with abiotic stressors [13] including cold, drought, and saline–alkali stress (Figure 1). Many crop plants that can withstand abiotic stress have previously been created via traditional marker-assisted breeding. However, due to extensive screening [34,35] and backcrossing procedures, it takes this tactic about a decade to generate abiotic stress-resilient crops effectively [36]. Although genetically modified, stress-tolerant plants have disclosed encouraging results, several barriers still stand in the way of their widespread commercialization. In many ways, crops with genome editing differ from genetically engineered species [37]. Considering this, genome editing seems to be a sophisticated strategy to create crops that are resistant to different abiotic stress in the future, because it allows precise manipulation of different gene loci in comparably less time, which lowers the cost of crop-improvement programmes [38]. Gene-editing technology based on CRISPR/Cas might successfully target complex quantitative genes linked either directly or indirectly to abiotic stressors. The use of CRISPR-Castechnology has been linked in recent years to the establishment of disease resistance in plants by modifying gene regulation [39,40,41,42]. Currently, CRISPR/Cas-based genome editing has been efficaciously utilized to investigate tolerance against multiple abiotic stresses, including heat, drought, salt, and nutritional values in several critical agricultural plants [43,44]. In this review article, we summarize the most likely uses of the CRISPR/Cas9-mediated genome editing technique in crop plants for dealing with diverse abiotic stresses such as heat, drought, salinity, cold, herbicide etc., and we predict the tools for future advancements in the creation of crop varieties that can withstand stresses.

2. Genome-Editing Strategy

Genome editing is one of the most promising approaches to understand the genome and to improve crop plants. The fundamental mechanisms involved in genetic modification by programmable nucleases (NHEJ) are the recognition of target genomic loci and binding of effector DNA-binding domain (DBD), double-stranded breaks (DSBs) in target DNA caused by restriction endonucleases (FokI and Cas), and repair of DSBs through homology-directed recombination (HDR) or non-homologous end joining [45]. While the well-organized and error-prone NHEJ results in the deletion or insertion of nucleotides, the less efficient and more accurate HDR results in the replacement of nucleotides. Genome-editing methods such as ZFN, TALEN, and CRISPR/Cas are being utilized to add the desired trait(s) and remove the undesirable ones. Numerous techniques are available for genome editing using either a site-specific recombinase (SSR) system or a site-specific nuclease (SSN) system. Both systems must be able to find a known sequence. The SSN system causes single or double strand DNA breaks and activates endogenous DNA repair systems. Depending on how the sites (loxP, FLP, etc.) are oriented, SSR technology, such as Cre/loxP- and Flp/FRT-mediated systems, can knockdown or knock in genes in the eukaryotic genome around the area of the target [46].
Plant genome-editing techniques have been classified into four major types based on onsite-specific endonucleases (Table 1). Those are ZFNs, meganucleases, TALENs, and CRISPR-Cas9 along with DSB-free genome editing, base editing, prime editing, and mobile CRISPR. These techniques are all discussed in detail below.

2.1. Zinc-Finger Nucleases

ZFNs are assemblages of DNA recognition modules based on zinc fingers and the DNA cleavage domain of the FokI restriction enzyme. With their use, the target genome can be altered to introduce a variety of genetic changes, such as deletions, insertions, inversions, translocations, and point mutations [47]. They have two domains, the first of which is a nuclease domain and the second of which is a DNA-binding domain. The DNA-binding domain’s 3- to 6-zinc finger repeats may recognize nucleotide sequences that are 9 to 18 bases long. The second domain is made up of the restriction enzyme Flavobacterium okeanokoites I (FokI), which is necessary for DNA cleavage [48].This method involves three artificial restriction enzymes, specifically ZFN-1, ZFN-2, and ZFN-3 [49]. ZFN-1: At this point, ZFN is transferred to the plant genome devoid of taking a repair template. Once it arrives at the plant genome, it makes double-stranded breaks (DSB) to the host DNA leading to non-homologous end joining (NHEJ) of DNA [50], which either produces site-specific arbitrary mutations or a small deletion or insertion. ZFN-2: Distinct from ZFN-I, a homology-directed repair (HDR) alongside a short repair template is delivered to the crop genome next to the ZFN enzyme [51]. The template DNA is homologous to the target DNA, which attaches to a specific sequence causing a double-stranded rupture. The template commences repairing with an endogenous repair mechanism which is directed to site-specific point mutations throughout homologous recombination (HR). ZFN-3: As soon as the ZFN transcribing gene is transferred to the plant genome next to the large repair template, it is called ZFN3 [51,52].
ZFN has been effectively implemented in Arabidopsis, tobacco, soybean, and maize [53,54,55,56]. In one example of the use of ZFNs in crop breeding, the insertion of PAT gene cassettes disrupted the endogenous ZmIPK1 gene in maize, which altered the inositol phosphate profile of growing maize seeds and improved herbicide resistance [53].ZFNs can be created utilizing various protein-engineering techniques to target essentially any unique DNA stretch [57]. ZFNs with enhanced specificity and activity have been developed to produce knockouts, which disable the gene’s function, as well as gain-of-function alterations [58].

2.2. Meganucleases

Longer DNA sequences (more than 12 bp) can be selectively detected and cut by meganucleases, which are endonucleases. This approach has been discovered in a wide variety of organisms, including archaebacteria, bacteria, algae, fungi, yeast, and many plant species. Meganucleases at the target region can sustain mild polymorphisms [59]. Meganucleases have been divided into five groups based on their sequence and structural features. These consist of His-Cys box, GIY-YIG, LAGLIDADG, PD-(D/E) XK, and HNH [60,61].Genome editing has mostly used members of the LAGLIDADG meganuclease (LMN) family. According to Silvaet al. [60], the name of this protein family is taken from the sequence of the main motif found in its structure. LMNs are typically expressed in the chloroplast and mitochondria of unicellular eukaryotes. The bulk of these endonucleases are dimeric proteins that have two separate functions: they splice their own introns as RNA maturases and cleave exon sequences as specialized endonucleases [62]. I-SceI and I-CreI’s genomes can be edited employing the rRNA gene of the mitochondrial DNA of Saccharomyces cerevisiae. The 21S contains the I-SceI gene’s location. The chloroplast of Chlamydomonas reinhardtii, a unicellular alga, was found to contain I-CreI, which is found in the 23S rRNA gene. However, due to the difficulties in reengineering meganucleases to target specific DNA areas, their utility in genome editing is limited [63].

2.3. Transcription Activator-like Effector Nucleases (TALENs)

Restriction enzymes called TALENs, or transcription activator-like effector nucleases, are designed to cleave specific DNA sequences. TALENs are made up of a nuclease that can cleave DNA in cells and a TALE domain that is intended to mimic the natural transcription activator-like effector proteins. Currently, a huge number of researchers are studying transcription activator-like effector nucleases (TALENs), which are composed of a free designable DNA-binding domain and a nuclease [64], in a variety of organisms. TALENs have recently emerged as a cutting-edge method for genome editing in a variety of species and cell types. It was discovered that TALENs may alter the genome in a variety of plants, including Arabidopsis, Nicotiana, Brachypodium, barley, potatoes, tomatoes, sugarcane, flax, rapeseed, soybean, rice, maize, and wheat [65,66]. According to a report, rice was the first crop in which TALENs technology was employed for enhancement. According to Li et al. [67], the main pathogen of blight disease (Xanthomonas oryzae) significantly reduces global rice production each year. By disrupting the genes for fatty acid desaturase (FAD), soybeans with high oleic acid and low linoleic acid levels were produced, improving the shelf life and heat stability of soybean oil [68,69]. TALENs are naturally occurring type III effector proteins created by Xanthomonas species that change the host plant’s gene expression. The TALENs proteins comprise a nuclear localization signal, a transcriptional activation domain, and a core DNA-binding domain [70]. The nuclear localization signal helps TALENs enter the nucleus, whilst the activation domain activates the transcriptional machinery to start expressing genes [71].

2.4. Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)/CRISPR-Associated Protein 9 (Cas9)

Clustered regularly interspaced short palindromic repeats (CRISPR/Cas9) are short, repetitive genetic variations that are present in most bacterial and archaeal species. CRISPR/Cas9 and its associated proteins produce a very strong defensive system that works as a safeguard for plants against foreign agents including bacteria, viruses, and other elements. The first application of CRISPR/Cas9 in an adaptive immune system was documented in a 2007 experiment [72]. The CRISPR/Cas9 gene-editing system has revolutionized research in animal and plant biology since its usage in genome editing was first demonstrated in mammalian cells in 2012 [73]. According to Rathore et al. [23] first-generation CRISPR/Cas9 genome editing involves simple manipulationand cloning techniques that can be applied to a variety of guide RNAs to edit different locations in the targeted organism’s genome (Figure 2). With the use of CRISPR/Cas, crop species can be precisely edited, opening the door to the generation of favorable germplasm and new, more sustainable agricultural systems. The genetic modification of crops can now be targeted and precise due to recent developments in CRISPR/Cas9 technology, hastening the advancement of agriculture [42]. To date, only a few species have been studied using this methodology [74].The yield, quality, disease resistance, and climatic adaptability of monocots and dicots have all been improved by the CRISPR/Cas9 system [75]. The genomes of cereal crops including wheat, maize, rice, and cotton as well as fruits and vegetables such as tomatoes and potatoes have all been altered using the CRISPR/Cas9 technique [76,77].
According to Makarova et al. [78], the CRISPR/Cas system can be divided into three types: type I, type II, and type III. Bacteria and archaea both have type I CRISPR/Cas mechanisms based on the exact signature of the Cas protein. The Cas3 protein’s endonuclease activity is used to connect to the DNA sequence [78]. In bacteria, the type II CRISPR/Cas system has been developed. The four protein pairs Cas1, Cas2, Cas4/Csn2 proteins, coupled with Cas9, make up the simplest system. The type III CRISPR/Cas system hunts for DNA and RNA in archaea, as well as infrequently in bacteria. Cas6, Cas10, and repeat associated mysterious proteins (RAMP) are markers for its presence. Cas10 protein’s processing of crRNA ultimately aims to cleave DNA [78]. The Streptococcus pyogenes (SpCas9)-derived type II CRISPR system mostly targets the negatively regulating genes [79].
The CRISPR/Cas technique is straightforward, stable, and enables effective change compared withthe first two generations of genome-editing systems. These traits allowed CRISPR/Cas to quickly replace the traditional genome-editing methods ZFN and TALEN. The techniquewas adapted from the bacterial defense mechanism. The CRISPR/Cas mechanism is used by a variety of bacterial and archaeal species to protect themselves against invading viruses [80]. Many studies are now being conducted to improve the CRISPR/Cas system and increase the tool’s ability to target the genome. For instance, non-canonical NGA and NG PAM sites in plants may be found using xCas9, SpCas9-VRQR, and Cas9-NG variants [81,82]. SpCas9 orthologues have been recognized from Streptococcus thermophiles (St1Cas9), Staphylococcus aureus (SaCas9), Streptococcus canis (ScCas9), and Brevibacillus laterosporus (BlatCas9).They have been demonstrated to amend plant genomic loci with PAM sequences of NNGRRT, NNG, NNAG AAW, and NNNCND, respectively [83,84]. Additionally, the type V Cas12a and Cas12b extracted from different bacterialsystems have been demonstrated with AT-rich PAM specifications and employed in genome editing of selected plants [85,86].
The CRISPR/Cas9 gene-editing approach has so far been used on more than 20 crop species to increase yields and reduce biotic and abiotic stress [87]. Genome-editing techniques based on CRISPR/Cas9 have been utilized to enhance agricultural disease resistance and tolerance to severe abiotic environments including salinity and drought. Three rice genes involved in regulating responses to various abiotic stress stimuli, including phytoene desaturase (OsPDS), betaine aldehyde dehydrogenase (OsBADH2), and mitogen-activated protein kinase (OsMPK2), have undergone sequence-specific CRISPR/Cas9-mediated genomic modification. CRISPR/Cas9 technology was successfully used by Shan et al. [88] to insert the TaMLO gene (mildew resistance locus O) into wheat protoplasts. It was also discovered that Blumeria graminis f. sp. Tritici, the agent of powdery mildew illness, is resistant to the CRISPR TaMLO knockdown (Btg). Wheat ethylene responsive factor3 (TaERF3) and wheat dehydration response element binding protein 2 (TaDREB2) are two abiotic stress-related genes that were targeted by the CRISPR/Cas9 genome-editing technology in wheat protoplasts, according to Kim et al. [89]. The CRISPR/Cas9 technology can be used in conjunction with current and upcoming breeding techniques such as speed breeding and omics-assisted breeding to boost agricultural production and ensure food security (Table 2).
Table 1. Comparison of different types of plant genome-editing techniques.
Table 1. Comparison of different types of plant genome-editing techniques.
FeatureZFNsMeganucleasesTALENsCRISPR/CasReferences
Length of target sequence (bp)18–36 bp12–40 bp28–40 bp20–22 bp[90,91]
Nuclease proteinFokII-SceIFokICas9 proteins[91,92,93]
DimerizationRequiredNot requiredNot requiredNot required[90,91,92]
Mode of actionDouble-stranded break in target DNADirect conversions in targeted regionsDouble-stranded break in target DNADouble-stranded breaks or single-stranded nicks in target DNA[94,95,96]
Repair eventsNHEJHDRHDRNHEJ[92,93,97]
MutagenesisHighMiddleMiddleLower[94]
CloningNecessaryNot necessaryNecessaryNot necessary[91,98,99]
Creation of libraries and multiplexingChallengingChallengingChallengingPossible[91,96,99]
CostHigherHigherHigherLow[100]
TypesOneOneOneMany[101]
SpecificityModerateHighHighLow[90,91]
Crop improvementLowLowLowHigh[100]
Future useMediumMediumMediumHigh[100]
Table 2. List of reported targeted gene(s) via ZFNs, TALEN, and MNs gene-editing tool technologies in different plant species to develop resistant/tolerant genotypes.
Table 2. List of reported targeted gene(s) via ZFNs, TALEN, and MNs gene-editing tool technologies in different plant species to develop resistant/tolerant genotypes.
CropGeneTraitTechniqueReferences
RiceOsQQRDetection of safe harbor loci herbicideZFNs[102]
OsBADH2, OsDEP1, OsSD1, OsCKX2FragranceTALEN[103]
Os11N3Bacterial blight resistanceTALEN[67]
OsCSAPhotoperiod sensitive male sterilityTALEN[104]
OsDERF1Drought toleranceTALEN[104]
WheatTaMLO-A1, TaMLO-B1, TaMLO-D1Resistance to powdery mildewTALEN[105]
MaizePATHerbicide resistanceZFNs[106]
ZmIPK1Herbicide tolerant and phytate reduced maizeZFNs[53]
ZmTLPTrait stackingZFNs[107]
ZmPDS, ZmIPK1A, ZmIPK, ZmMRP4Biosynthesis of phytic acidTALEN[108]
MS26Independent lines of male sterile plantsMNs[109]
BarleyHvPAPhyPhytase reduction and seed developmentTALEN[110]
SoybeanDCLHerbicide transmissionZFNs[111]
FAD2-1A, FAD2-1BLow polyunsaturated fats TALEN[68,69]
TobaccoGUS: NPTIIChromosome breaksZFNs[112]
Endochitinase-50 gene (CHN50)Emergence of resistance to herbicidesZFNs[113]
TomatoL1L4/NF-YB6Reduced contents of the anti-nutrient’s oxalic acidZFNs[114]
CottonEPSPSHerbicide toleranceMNs[115]
HppdHerbicide toleranceMNs[115]
PotatoVInvSugar metabolismTALEN[116]

2.5. DSB-Free Genome Editing

A sole histidine residue at site 840 of the HNH domain of SpCas9 cuts the PAM strand, while the aspartate at site 10 in the RuvC domain cuts the opposite strand3. Mutating both amino acids to alanines (D10A and H840A) resulted in nuclease-dead Cas9 (dCas9). dCas9 still identifies its target site and frees up the DNA in an R-loop without including DSBs. The binding of dCas9 to its solitary target site can work as a repressor of transcription and is called CRISPR interference (CRISPRi). Alternately, dCas9 can be utilized as a tool for localization of DNA effector proteins to the genome. Examples of this approach are CRISPR–DNMT3 fusion proteins and CRISPR activators (CRISPRa) for targeted methylation. DNA-alteration enzymes are combined with dCas9 to induce genetic variants for overcoming the limitations linked with DSB initiation in genome engineering [117].

2.6. Base Editing

The first base editor combines dCas9 to the cytidine deaminase apolipoprotein B mRNA editing catalytic polypeptide-like (rAPOBEC1), which catalyzes the alteration from cytidine to uracil. The cell mends this uracil into thymidine, resultingin an assembly (BE1) replacing a C•G by a T•A base pair, entitled a cytosine base editor (CBE) [118]. First-generation CBEs were suppressed by uracil glycosylation. So, second-generation base editors (BE2) were invented by combining an uracil glycosylase inhibitor (UGI) with the dCas9–rAPOBEC1 combination [119].For increasing editing efficiency, dCas9 can be changed into a nickase SpCas9-D10A (BE3). The strand not altered by rAPOBEC1 is cleaved. The cell identifies this nick and starts DNA repair to solve the damage. The strand withthe base modification is used as a template for repairing the nick to yield stable integration. The BE3 architecture was furthermore ameliorated by combining an additional UGI in fusion with linker optimization to result in a fourth-generation cytosine base editor (BE4). BE4s have improved editing efficiency by approximately50%, with two-fold decline of unintended byproduct formation such as point mutations and indels [118]. Subsequent ancestral reconstitution and codon optimization led to a CBE architecture that enables the most powerful base editing in organoids, 2D cell lines, and in vivo by improving nuclear localization and expression of the proteins [120].

2.7. Prime Editing

The logic behind prime editing is to escort exogenous DNA with the modification of interest close to the Cas9 binding site. Areverse transcription (RT) domain obtained from the Moloney murine leukaemia virus was combinedwith nickase SpCas9- H840Atodevelopthe first generation of prime editors (PE1). The RT domain changes RNA into DNA tofind its template in the 3′ extension of the specially designed sgRNA, entitledthe primeediting guide RNA (pegRNA).Itguides the Cas9 in PE1 to the target site. After targetrecogination, the PAM-consistingstrand is nicked by the active HNH domain of Cas9-H840A. Then, the pegRNA extension combineswiththe nicked strand of the primer-binding site (PBS).Then, the RT domain of PE1 uses the restpegRNA(RT template) to synthesize a 3′-DNA flap containingthe edit of interest. This DNAflap is solved by cellular DNA repair procedure combining the edit of interest [121]. Theprime editing requires optimizing PE3guides andpegRNA, limiting its implementationin organoids. Threemodifications have been made forovercoming this issue. First, the utilizationof two pegRNAs in trans alongwith overlayingRT domains enhancesprime-editing competencein plants [121]. Second, engineered pegRNAs can have tmpknot or evopreqdomains combinedatthe 3′ end. These domains enhancethe stability of the pegRNA [122]. Finally, including the N394Kand R221K amino acid alterationincreases the nuclease workof SpCas9, resulting in a more efficient PE2Max [123].

2.8. Mobile CRISPR

A breakthrough in the CRISPR tool, “genetic scissors” was announced by scientists of the Max Planck Institute of Molecular Plant Physiology to edit plant genomes. The discovery could speed up and simplify development of novel and genetically stable crop varieties by fusing grafting with a ‘mobile’ CRISPR tool. The drawing of the CRISPR/Cas9 gene scissors is transferred as RNA from the rootstock of a genetically modified plant to the grafted shoot of a normal plant. The gene scissors protein is made with the aid of the RNA. This gene scissor protein edits specific genes in flowers. Plants carry the desired gene modification in the next generation. A normal shoot is grafted onto roots containing a mobile CRISPR/Cas9, which allows the genetic scissor to move from the root into the shoot. It edits the plant DNA without leaving a trace of itself in the subsequent generations of plants. This ground-breaking turn can save cost and time and evade current limitations of plant breeding.

3. Genome Editing Related to Abiotic Stresses

Abiotic stresses that impact plant growth and development, such as salt, drought, extremely high temperatures, cold, and heavy metals, can reduce agricultural production by approximately 50% [124].Numerous biochemical, morphological, and physiological factors important for plant development are influenced by stress. Stresses from the environment can modify how plants behave as they develop. Most changes in plant growth and development caused by different abiotic stresses are associated with poorer yields [13]. By 2050, the rapid growth in the human population is predicted to reach 9.7 billion. The global temperature is also set to increase significantly. As plant scientists, it is hard for us to manage the food requirements of the increasing population. However, we own the capability to develop climate-flexible crop varieties that can flourish under such challenging circumstances. These varieties must be maintained in ruthless climatic conditions such as heat, drought, heavy metals, cold, or flood stresses. This requires a continuous search for newer and diverse germplasm [125,126], which was traditionally performed either entirely through development of natural variations [127,128] or by selective breeding [129,130]. Another possibility is the construction of mutant populations that are evaluated to hunt for new resources among variations that might be novel valuable mutations that in turn are included in breeding programmes. Modern genome-editing system tools such as CRISPR facilitate the user to commence desirable genomic modifications accurately, illustrating great promise as a tool for producing novel climate-resistant plants [131]. In over 20 agronomically important crops, CRISPR/Cas mediated gene editing is widely utilized and accepted for crop improvement against different abiotic stresses [79].
Ordinarily, plants are equipped with numerous defense schemes against abiotic stresses. Among numerous defense mechanisms of abiotic stresses, the five broad-spectrum protections are regulated utilized in a complicated managing network consisting of numerous mediators and gene regulatory constituents in response to abiotic stresses [132]. During the procedure, stress hormones, particularly nitrogen oxides (NO), abscisic acid (ABA), polyamines (PAs), calcium ions (Ca2+), hydrogen sulfide (H2S), reactive oxygen species (ROS), and phytochrome B (PHYB), interact with others, either synergistically or antagonistically. The transcription factors (TFs) could alter the expression of genes and enzyme activity in a regulatory way, triggering a suitable reaction. The regulatory constituents open a lot of potential for developing multiple stress tolerance/resistance. Five main plant defenses to abiotic stresses are ROS scavengers, molecular chaperones, cuticle as the outer shield, oxylipin precursors, and osmoprotectants, along with unsaturated fatty acids, and compatible solutes [132].

3.1. Drought Stress

Drought is becoming a challenge to sustainable agriculture due to the consequences of climate change, including erratic rainfall patterns and rising temperatures in many regions of the world. The greatest danger to global food security is drought stress, which is the primary factor in the catastrophic loss of agricultural production and productivity [133]. Drought alone can reduce yield by 50–70% in different crops [134]. For example, 40% yield losses due to drought stress have been reported in maize [35,135], 50% in rice [136], 21% in wheat [126,135], 27–40% in chickpea [125,137], 68% in cowpea [138] and 42% in soybean [34,139]. After the discovery of genome editing, efforts are being planned to alter the genes involved in pathways enabling drought tolerance, in order to increase farmers’ acceptance of crops using these technologies. In recent years, in-depth research has helped to adapt and overcome drought stress using CRISPR-Cas9 technology (Table 3).
In many crop plants, H2O2 and abscisic acid (ABA) are frequently produced in situations of salinity or drought stress. The discovery was reported of ABA-induced transcription repressors (AITRs) as a novel transcription factor family that plays a significant role as feedback regulators of ABA signaling. Alternation in the expression of AITR genes resulted in abiotic stress tolerance, including drought and salinity in Arabidopsis [140,141]. A CRISPR/Cas9-induced mutation in the Arabidopsis OST2 structural gene exhibited drought resistance [142]. Another study found that knockout of Arabidopsis plants’ genemiR169athrough CRISPR/Cas9 led to significantly improved drought tolerance [143]. Similarly, Arabidopsis’ drought tolerance increased after the vacuolar H+-pyrophosphate (AVP1) regulating gene was expressed using CRISPR/Cas9 [144]. Similar results were shown when the abscisic acid-responsive element binding gene (AREB1) was activated in Arabidopsis through CRISPR/Cas9a [145]. Recently, drought tolerance in Arabidopsis thaliana was demonstrated via the CRISPR/Cas9 gene silencing of the trehalose (TRE1) gene [146].
Numerous studies have documented how CRISPR confers drought resistance in many plants. For instance, it has been demonstrated that increasing rice’s ability to withstand drought can be attained by reducing the expression of the regulatory genes DERF1, PMS3, MSH1, MYB5, and SPP [147]. In rice plants, drought stress tolerance increased after OsERA1 was modified using CRISPR/Cas9 [148]. CRISPR/Cas9 has been employed to improve drought resistance in rice by knocking out the SRL1, SRL2, and ERA1 genes [148,149]. A CRISPR/Cas9-created ospyl9 mutant might increase rice yield and drought tolerance [150]. Indica mega rice cultivar MTU1010 with broader leaves, a decreased stomatal density, and improved leaf water retention under drought stress was developed using CRISPR/Cas9 to modify the OsDST gene [151]. The OsOREB1, OsRab21, OsRab16b, OsLEA3, OsbZIP23, OsSLAC1, and OsSLAC7 genes, which act downstream of SAPK2, were modulated in expression in the loss-of-function sapk2 mutant of rice plants developed using CRISPR/Cas, increasing their tolerance to drought stress [131].
Two genes, RVE7 and 4CL, have been found to be associated with drought tolerance in chickpeas. The first report of CRISPR/Cas9-mediatedediting of the chickpea protoplast was made by Badhan et al. [152]. They described knockouts of the genes 4CL and RVE7, which are linked to pathways for drought tolerance. That study established a framework for potential future chickpea-genome-editing approaches [153]. Another gene, namely ARGOS8, responding to drought stress has been altered through genome editing. The expression of the ARGOS8 gene increased as a result of negative regulators of ethylene signaling pathways, providing drought tolerance [154,155]. To increase the production of maize under drought stress under field conditions, the GOS2 promoter region was replaced with an ARGOS8 promoter sequence using the CRISPR/Cas system [156].
CRISPR/Cas9 altered the GID1 gene in tomato plants, which exhibit high leaf water content under drought conditions [157]. Additionally, SlLBD40 gene mutation caused by CRISPR/Cas9 significantly improved drought tolerance in tomato [158]. Furthermore, use of the CRISPR/Cas technique to alter mitogen-activated protein kinases (MAPKs) revealed SlMAPK3 to be a drought stress modulator [159]. Knockout of the SINPR1 gene resulted in increased drought tolerance and down-regulation of drought-related genes [160].
Drought resistance of wheat was improved by CRISPR/Cas editing of wheat TaDREB2 and TaERF3 [89]. In wheat, a multiplex CRISPR/Cas9 assay was used to alter the SAL1 gene, a negative regulator of drought tolerance, to increase drought tolerance at the seedling stage [161]. CRISPR/Cas genome editing of the HB12 gene can increase cotton’s resistance to drought [162]. CRISPR/Cas9 was used to modify the BnaA6.RGA gene in oil seed crops, which significantly improved rapeseed’s ability to withstand drought [163].

3.2. Heat/Temperature Stress

Plants have a preferred temperature, any rise or fall in that temperature can significantly impede their development and productivity. The third most important abiotic factor is heating stress, which may decrease crop production considerably. For instance, every 1 °C augmentation in atmospheric temperature diminishes wheat yield by 6%, rice yield by 10–20%, and corn yield by 21–31% [164,165,166]. Significant yield losses were caused by high heat stress, which is now recognized as a severe problem that will simply become worse in the future. All phases of plant growth, from germination to harvest, are severely harmed by heat stress [167,168]. Heat stress not only increases plant mortality rates but also reduces plant quality [169,170].
In severe cases, a bad alteration in temperature results in plant mortality because plants are more susceptible to temperature changes. The ideal temperature would normally be better for crop growth and development; conditions below and above the optimum temperature have a harmful effect on productivity. For every 10 °C rise, followed by 20 °C and 30 °C, mostbiochemical and enzymatic procedures double in speed [171]. Abiotic stressors, predominantly high and low heat, have a harmful effect on the premature stage of the male gametophyte in a range of agricultural crops, including maize, rice, barley, wheat, sorghum, and chickpea [172]. Due to temperature stress, the functions of tapetal cells are diminishedduring the reproductive growth period, and the anther is dysplastic. Pollen discharge is insufficient and indehiscence happens as a result of increased heat preventing pollen grains from escalating. Plants have developed precise physiological and chemical reactions to manage temperature stress [173].
The presence of genes that are responsive to heat stress, signal transduction, and the synthesis of metabolites are only a few of the complex molecular systems that plants activate in response to heat stress. Different temperature-stress-related genes have been identified and characterized to improve plants’ ability to withstand heat as a result of developments in structural and functional genomics technologies in plants. The heat stress reaction, which is connected to the accumulation of ROS, is mediated by the heat shock transcription factors (HSFs) and the heat shock proteins (HSPs) [174].Therefore, by enhancing plants’ ability to resist ROS components, temperature stress tolerance can be improved [175]. This indicated that higher tolerance might increase the antioxidant properties of crops. Plant temperature tolerance was significantly increased via metabolite production and temperature-induced gene expression. To explore the molecular processes associated with temperature stress and improve plant heat tolerance, CRISPR-Cas9 is a cutting-edge technology among all genome-editing techniques [176] (Table 3).
A cultivable HS-inducible rice mutant was created using CRISPR/Cas9 technology [177]. The orthologs of mitogen-activated protein kinase 3 and agamous-like 6 were modified using CRISPR to increase tomato sensitivity to heat stress, whereas ADP-ribosylation factor 4 enhanced tomato sensitivity to salinity shocks. According to Bouzroud et al. [178], these CRISPR-edited mutant plants had improved agronomic characteristics and were resilient to abiotic stresses. As a component for heat tolerance, BRZ1 positively regulates the formation of ROS in the tomato apoplastic area. This was confirmed by the CRISPR-Cas9-based bzr1 mutants, which showed reduced temperature tolerance and respiratory burst oxidase homolog 1 (RBOH1) with diminished hydrogen peroxide generation in the apoplast [179]. In comparison to wild-type crops, the development of CRISPR/Cas-mediated heat-stress-sensitive albino 1 (HSA1) mutants of tomato showed greater sensitivity to temperature stress [180].
The thermosensitive genic male sterile gene was altered by CRISPR in maize to promote thermo susceptible male-sterile plants [181]. In lettuce, knockouts of NCED4, a crucial regulating enzyme in abscisic acid production, allowed the seeds to germinate at a higher temperature. As a result, LsNCED4 mutants may have commercial significance in manufacturing environments with high temperatures [182]. In order to make a plant more resistant to heat, the hsps gene, which increases osmolyte levels and prevents cell protein damage, can be overexpressed [183]. The protein kinase SAPK6 and the transcription factor OsbZIP46CA1 in rice also increase the capacity for responding to heat stress [184].

3.3. Cold Stress

Cold stress, which includes chilling (20 °C) and freezing (0 °C) temperatures, hinders plant growth and development and severely limits plant geographic expansion and agricultural productivity [185]. Plants are directly inhibited from responding metabolically to low temperatures, which results in osmotic stress, oxidative stress, and other types of stress. Due to mechanical damage and metabolic dysfunction caused by extreme cold temperatures, plant growth and development are halted [186]. The physiological, biochemical, and molecular behavior of plants during their growth and expansion is adversely affected by cold stressors. The photosynthetic capacity and crop anatomy are brutally impacted by cold exposure, especially throughout the winter [187,188].Cold stress during the seedling stage may cause impaired germination and emergence. Long-term exposure impairs source–sink relationships, growth, nutrient localization, and leaf chlorosis [189]. Membrane formation, which amplifies other cold-stress-related downstream processes, is the main consequence of cold stress on crops [190]. In-generic or inter-specific hybridization has been successful in boosting the cold tolerance of significant crops using conventional breeding methods. For creating non-transgenic genome-edited crops to combat climate change and ensure future food security, CRISPR/Cas9 is a clever and practical approach [191,192] (Table 4).
To increase the plant’s resistance to cold, genome editing is employed to target a few of the depressant regulator transcription factors in rice. A transcription factor called OsMYB30 attaches to the amylase gene promoter and negatively affects cold tolerance. According to Lv et al. [193], under conditions of cold stress, OsMYB30 forms a compound with OsJAZ9 and slows down the expression of the amylase gene, which may contribute to increasing cold sensitivity by causing maltose buildup and starch breakdown. In order to determine the specific function of the TIFY1a, TIFY1b, and Ann3 genes in rice’s ability to withstand cold stress, CRISPR/Cas9 technology has also been applied to these genes. The mutant outperformed the natural variation in terms of yield, temperature tolerance, and amount of germination prior to harvest [194]. Using CRISPR base editing, suppression of photosynthetic genes in rice plants under cold stress has been shown to cause the white-striped leaves phenotype in the white stripe leaf 5 (wsl5) mutant line [195,196].
PRPs are proline-rich proteins that not only aid in dealing with low temperatures but also reduce nutrient loss, boost antioxidant activity, and aid in the production of chlorophyll. Rice capacity for cold tolerance was improved by the CRISPR/Cas9 deletion of OsPRP1, which encodes a proline-rich protein [197]. In a recent work using CRISPR/Cas9, three rice genes, viz., OsPIN5b, GS3, and OsMYB30were altered to increase spike length, grain size, and resilience to cold stress [198]. The CRISPR/Cas9 technology altered the G-complex-related genes i.e., OsRGA1, OsGS3, OsDEP1, and OsPXLG4 to make rice more resistant to chilling stress [199].Because tomato plants are prone to chilling stress, their fruits are more vulnerable to damage from the cold. C-repeat binding factor 1 (CBF1) was shown using CRISPR-Cas9-based cbf1 mutants to protect the tomato plant next to it from cold/chilling damage and decrease electrolyte leakage [200]. These plants also demonstrated excellent addition of hydrogen peroxide and indole acetic acid, resulting in tomato plants tolerant of chilling stress.
Table 3. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for development of tolerant genotypes against drought and heat stresses.
Table 3. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for development of tolerant genotypes against drought and heat stresses.
CropsGeneTraitTechniqueReferences
RiceOsDERF1DroughtCRISPR/Cas9[147]
RiceSRL1, SRL2DroughtCRISPR/Cas9[149]
RiceOsAAA-1, OsAAA-2DroughtCRISPR/Cas9[201]
RiceOsNAC006 (transcription factor)Drought and heat sensitivityCRISPR/Cas9[202]
RiceOsAOX1aDrought resistanceCRISPR/Cas9[147]
RiceOsDSTDrought and salinityCRISPR/Cas9[151]
RiceOsERA1, OsPYL9DroughtCRISPR/Cas9[148,150]
RiceSAPK2Tolerance to salinity and droughtCRISPR/Cas9[131]
RiceOsPMS3Photoperiod-sensitive male-sterileCRISPR/Cas9[147]
RiceCsaPhotosensitive-genic male-sterileCRISPR/Cas9[203,204]
RiceTMS5Thermo-sensitive genic male-sterileCRISPR/Cas9[205]
RiceOsNAC14Drought toleranceCRISPR/Cas9[206]
RiceOsPUB67Drought toleranceCRISPR/Cas9[207]
WheatTaDREB2, TaERF3Tolerance to droughtCRISPR/Cas9[89]
MaizeZmARGOS8DroughtCRISPR/Cas9[156]
MaizeZmTMS5Creation of thermosensitive maize linesCRISPR/Cas9[181]
MustardBnaA6.RGADrought toleranceCRISPR/Cas9[163]
SoybeanDrb2a, Drb2bTolerance to drought and salinity stressCRISPR/Cas9[208]
SoybeanGmMYB118Drought toleranceCRISPR/Cas9[209]
Chickpea4CL, RVE7Drought toleranceCRISPR/Cas9[152]
TomatoSIMAPK3 and SlNPR1DroughtCRISPR/Cas9[159,160]
TomatoSlARF4DroughtCRISPR/Cas9[140]
TomatoSIAGL6Heat stressCRISPR/Cas9[210]
Table 4. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for development of tolerant genotypes against cold stresses.
Table 4. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for development of tolerant genotypes against cold stresses.
CropsGeneTraitTechniqueReferences
RiceOsMYB30Cold toleranceCRISPR/Cas9[198]
RiceOsAnn3Cold toleranceCRISPR/Cas9[211]
RiceOsAnn5Cold toleranceCRISPR/Cas9[211]
RiceOsPRP1Cold tolerance CRISPR/Cas9[212]
TomatoSlCBF1Cold toleranceCRISPR/Cas9[200]
Arabidopsis thalianaAtCBF1, AtCBF2Cold toleranceCRISPR/Cas9[213]

3.4. Salinity Stress

Owing to the negative consequences of climate change, salinity stress has recently become much worse [214]. Salinity stress is the second most severe abiotic danger that affects fertile lands as well as crop productivity [215]. According to Morton et al. [216] and Van Zelm et al. [217], severe salts have an impact on about one-fifth of the irrigated agricultural area. Lack of good irrigation water, a changing climate, and excessive use of chemicals such as fertilizers and pesticides prolong the process of adding more land to the salinity stress zone. According to estimates made by Jamil et al. [218], 50% of cultivable lands will be saline by 2050 due to the overuse of chemicals including fertilizers and pesticides. One of the most important and harmful factors that has a negative impact on soil quality and agricultural output is salt stress. When too many soluble salts accumulate in the crop root zone, it causes salinization of the soil because roots are unable to absorb water. Thus, osmotic stress and nutritional imbalance in plants have a negative impact on their morphology, biochemistry, and biomass, which ultimately causes irreparable plant damage [219,220,221].
Reactive oxygen species (ROS) are intensified by salt stress, which has a detrimental effect on crops’ cellular and metabolic processes [222,223]. Lipid peroxidation, which causes membrane deterioration as well as protein and DNA damage, is a harmful effect of ROS [224]. By diminishing chlorophyll content and stomatal conductance, salt stress hinders the development of the photosystem II and the transpiratory apparatus [225]. Additionally, it decreases the water potential of the soil and leaves, which lowers plant turgor pressure by affecting water relations and causing osmotic stress [226]. Plants suffer from decreased leaf area, lower photosynthetic rate, poor seed germination, decreased biomass production, and crop yield as a result [227,228,229]. Salinity tolerance is the ability of a plant to maintain the equilibrium of biomass and/or output under conditions of salt stress. In order to tolerate salt, plants have several molecular and physiological mechanisms [230].
Genome editing has the capacity to improve crops; there are yet few studies on its effective application in breeding plants that can withstand saline stress (Table 5). In one such work, rice was modified to impart salt stress tolerance by editing the OsRR22 gene, which encodes for a transcription factor (TF) involved in the control of signaling and the metabolism of cytokinins in plants [231,232]. Using CRISPR/Cas9 technology, the OsRR22 gene was altered, and two homologous T2 generations revealed improved salt tolerance with no discernible difference between the modified and wild-type lines [232]. Using CRISPR/Cas9 technology, the paraquat tolerance-3 mutations (OsPQT3) gave rice a high level of salt tolerance [233]. The function of OsmiR535 in salt stress tolerance was investigated using genome-editing techniques, and it was proposed that OsmiR535 might be knocked out using CRISPR/Cas9 to enhance salinity tolerance in rice. Additionally, a homozygous 5bp deletion in the OsmiR535 coding region might be a valid target for raising rice’s salt tolerance [234]. Furthermore, some other genes increase the ability of rice to tolerate salt, using CRISPR/Cas9 technology by eliminating the OsbHLH024 gene and increasing the expression of the ion transporter genes including OsHKT1;3, OsHAK7, and OsSOS1 [235]. When the rice OsRAV2 gene was altered using CRISPR-Cas, the rice plants were able to survive under high salt conditions [236].
Improvements in salt stress tolerance were seen in tomatoes after changes were made to the 8CM and PRD domains of the hybrid proline-rich protein1 (HyPRP1) [247]. Additionally, the capability of crops to tolerate salt stress may be significantly increased by employing CRISPR/Cas9 technology to eliminate the OsDST genes for rice [151], OsNAC041 [238], and HvITPK1 [246] for barley.

3.5. Heavy Metals Stress

An important issue for sustainable agricultural development is heavy metals, which seriously impair plant growth and productivity [249]. Heavy metals (HMs) including Mn, Cu, Ni, Co, Cd, Fe, Zn, and Hg, among others, have accumulated in soils as a result of various human activities such the application of fertilizer, incorrect disposal of industrial waste, and unauthorized sewage disposal [250,251], or the hasty disposal of vehicle waste. They are either collected on the soil surface or leached from the soil into the groundwater [252,253]. Additionally, heavy metals cause oxidative stress by promoting the generation of hydroxyl radicals (OH), superoxide radicals, and hydrogen peroxide (H2O2) [250,254]. Plant physio-morphological activities are hampered by the accumulation of HMs, especially in the roots where they are blocked by Casparian strips or trapped by root cell walls, which eventually reduces crop output [255]. When consumed, heavy metals accumulated in plants canseriously impair human health [256].
To combat heavy metal stress in plants, CRISPR-Cas9-induced plant mutants may prove useful (Table 6). In contrast to WT Co10 plants, the oxp1/CRISPR mutant of Arabidopsis plants exhibits resistance to Cd, indicating an increased capacity for heavy metal detoxification in mutant crops [257]. Accordingly, study showed how indel mutations using gene-editing techniques could provide tolerance to heavy metals and xenobiotics in plants [257]. Increased plant tolerance to heavy metals is influenced by a variety of genes [258]. Several transporter genes in rice, including OsLCT1 and OsNramp5, are implicated in Cd absorption by the roots [259]. The amount of Cd in rice has been reduced by CRISPR/Cas9-enabled gene-expression manipulation. Rice grains with OsNRAMP1 knocked out by CRISPR/Cas9 have decreased levels of Cd and lead (Pb) [260,261]. Eliminating an R2R3 MYB transcription factor called OsARM1 using CRISPR/Cas9 prevents rice from absorbing and transporting arsenic [262].Cesium (Cs+) absorption and translocation in rice are regulated by the OsHAK1 gene. Using the CRISPR-Cas9 technique, the cesium permeable potassium transporter OsHAK1 was turned inactive [263].

3.6. Herbicide Stress

In order to increase crop productivity, there is a need to manage weed growth with application of herbicides. Herbicides destroy non-target plants while also causing stress to the target plants and weed plants by interfering with or changing their metabolic processes. They also leave soil residues that are hazardous to the environment [264,265].The morphological, physiological, and biochemical traits of agricultural plants have been negatively impacted by the inappropriate application of herbicides. Herbicide toxicity reduces photosynthetic activity, which has a detrimental impact on the ability of crop plants to produce yield. One of the main goals for raising agricultural productivity is the development of herbicide tolerance in crop plants. To improve herbicide resistance in plants, genome editing including ZFNs, TALENs, and CRISPR/Cas technologies is an excellent tool (Table 6).
Leucine, isoleucine, and valine are branched amino acids whose biosynthesis is catalyzed by the enzyme acetolactate synthase, which is encoded by the ACETOLACTATE SYNTHASE (ALS) gene [266,267]. It is a potential target of many herbicide improvement programmes. The recombination of acetolactate synthase using CRISPR/Cas9 produces herbicide resistance in rice [268] and in watermelons [269]. Additionally, using the same strategy and emphasizing the ALS1 and ALS2 genes, herbicide-resistant maize plants were produced [270]. CRISPR-based editing in the OsALS1 gene has been used to introduce herbicide tolerance characteristics into rice [271,272]. Glyphosate is one of the most imperative and quickly adopted herbicides for function in resistant crops such as soybean, maize, sugar beet, and chili pepper. The advancement of glyphosate-resistant plants requires changes in the machinery of some genes [203]. 5-enolpyruvylshikimate-3-phosphate synthase (EPSPS) enzyme is implicated in the formation of aromatic compounds in crops with the transfer of phosphoenolpyruvate (PEP) enzyme for activating the reaction [203,273]. Glyphosate hinders the act of the EPSPS enzyme by inhibiting the add-on of glyphosate to the PEP enzyme binding sites, eventually blocking the formation of aromatic products and causing crop death [203]. The endogenous EPSPS gene of rice was targeted with CRISPR/Cas9 to produce site-specific gene incorporation and substitution, which were fully transferred to the next generation with crops 100% resistant to the glyphosate [203]. CRISPR/Cas9 was also utilized toproduce a mutation in the promoter of the EPSPS gene of chili to state this gene beneath the action of glyphosate [274]. The resulting crops were reasonably resistant to glyphosate, and additional studies advised that selecting a diverse promoter may assist in the development of entirely resistant chili [274].The modified genotypes of rice and flax now have enhanced tolerance to glyphosate as a result of the CRISPR/Cas9 change of two nucleic acid residues in the binding site of glyphosate–EPSPS [91,203]. Recently, herbicide resistance was developed in tomato plants by CRISPR-Cas9-based targeted mutations in EPSPS, PDS (phytoene desaturase), and ALS [92].
Table 6. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for tailoring herbicide and metal stress tolerance.
Table 6. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for tailoring herbicide and metal stress tolerance.
Crops GeneTraitTechniqueReferences
RiceC287THerbicide resistanceCRISPR/Cas9[274]
RiceBELHerbicide resistanceCRISPR/Cas9[71]
RiceOsALS1Herbicide toleranceCRISPR/Cas9[271]
RiceEPSPSHerbicide resistanceCRISPR/Cas9[203]
RiceSF3B1Herbicide resistanceCRISPR/Cas9[72]
WheatALSHerbicide resistanceCRISPR/Cas9[275,276]
MaizeALS1 and ALS2Herbicide resistanceCRISPR/Cas9[270]
MaizeMS26Herbicide resistanceCRISPR/Cas9[270]
SoybeanALS1Resistant to ChlorsulfuronCRISPR/Cas9[277]
TomatoALSResistant to ChlorsulfuronCRISPR/Cas9[278]
TomatoSlEPSPSHerbicide resistanceCRISPR/Cas9[92]
TomatoSlALS1, SlALS2Herbicide resistanceCRISPR/Cas9[92]
TomatoSlpds1Herbicide resistanceCRISPR/Cas9[92]
RiceOsTubA2Base editingCRISPR/Cas9[279]
RiceOsHAK1Low cesium accumulationCRISPR/Cas9[263]
RiceOsPRX2Potassium deficiency toleranceCRISPR/Cas9[280]
RiceOsARM1Increase tolerance to higharsenicCRISPR/Cas9[260]
RiceOsLCT1Less cadmium accumulationCRISPR/Cas9[259]

4. Conclusions and Prospects

Plants serve as sources of food, fiber, medicine, biofuels, and other goods. Farmers need new, superior cultivars in order to increase crop output and feed both the nation and the world. Plant breeders need a variety of tools for this purpose, including genomics and marker-assisted molecular breeding. Scientists can now implant desired traits more precisely and faster than in the past. Meganucleases (MNs), zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and the clustered regularly interspaced short palindromic repeats (CRISPR) system are genome-editing tools that have been used with greater accuracy and efficiency than conventional breeding to enhance the quality of staple, oilseed, and horticultural crops. Today, there are several successful cases of “genome editing.” In order to edit genes accurately in the genomes of model and crop plants as well as a range of other organisms, genome editing employs designed nucleases as potent tools that target certain DNA sequences. A study of the literature on transcriptomics, biotechnology, genomics, and phonemics has shown that this novel approach to crop development is effective. CRISPR/Cas9-based genome editing is a genuinely innovative strategy. With genome editing, crops can effectively incorporate a variety of genetic traits. When these precise and powerful methods are applied to expedite plant breeding, they create certain outcomes. In order to accomplish a second Green Revolution and meet the escalating food demands of a quickly growing global population under constantly changing climatic conditions, plant breeding will advance with the help of this multidisciplinary approach. By overcoming the limitations of current transgenic techniques, genome-editing technology ushers in a new era of improved plant genetics. This information may be proved useful to plant breeders and researchers in their thorough evaluation of the use of various gene-editing tools to improve crops by focusing on the targeted gene. We believe that CRISPR/Cas9 technology islikely to bridge the GMO and societal divide in upcoming days.

Author Contributions

Writing—original draft preparation, R.K.Y., S.T., N.T., R.A., S.C., P.N.T. and D.K.P.; writing—review and editing, M.K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohidem, N.A.J.; Hashim, N.; Shamsudin, R.; Che Man, H. Rice for Food Security: Revisiting Its Production, Diversity, Rice Milling Process and Nutrient Content. Agriculture 2022, 12, 741. [Google Scholar] [CrossRef]
  2. Jaganathan, D.; Ramasamy, K.; Sellamuthu, G.; Jayabalan, S.; Venkataraman, G. CRISPR for Crop Improvement: An Update Review. Front. Plant Sci. 2018, 9, 985. [Google Scholar] [CrossRef] [PubMed]
  3. Mahood, E.H.; Kruse, L.H.; Moghe, G.D. Machine Learning: A Powerful Tool for Gene Function Prediction in Plants. Appl. Plant Sci. 2020, 8, e11376. [Google Scholar] [CrossRef]
  4. Parkhi, V.; Bhattacharya, A.; Choudhary, S.; Pathak, R.; Gawade, V.; Palan, B.; Alamalakala, L.; Mikkilineni, V.; Char, B. Demonstration of CRISPR-cas9-mediated Pds Gene Editing in a Tomato Hybrid Parental Line. Ind. J. Gen. Plnt. Breed. 2018, 78, 132–137. [Google Scholar] [CrossRef]
  5. Ni, Z.; Han, Q.; He, Y.-Q.; Huang, S. Application of Genome-Editing Technology in Crop Improvement. Cereal Chem. 2018, 95, 35–48. [Google Scholar] [CrossRef] [Green Version]
  6. Rothstein, R.J. One-Step Gene Disruption in Yeast. Methods Enzymol. 1983, 101, 202–211. [Google Scholar] [CrossRef]
  7. Sharma, P.; Tiwari, S.; Tripathi, N.; Mehta, A.K. Polymorphism Analysis in Advanced Mutant Population of Oat (Avena sativa L.) Using ISSR Markers. Physiol. Mol. Biol. Plants 2016, 22, 115–120. [Google Scholar] [CrossRef] [Green Version]
  8. Thomas, K.R.; Folger, K.R.; Capecchi, M.R. High Frequency Targeting of Genes to Specific Sites in the Mammalian Genome. Cell 1986, 44, 419–428. [Google Scholar] [CrossRef]
  9. Brookes, G. Genetically Modified (GM) Crop Use 1996–2020: Environmental Impacts Associated with Pesticide Use Change. GM Crops Food 2022, 13, 262–289. [Google Scholar] [CrossRef]
  10. Tripathi, M.K.; Tripathi, N.; Tiwari, S.; Mishra, N.; Sharma, A.; Tiwari, S.; Singh, S. Identification of Indian Soybean (Glycine max [L.] Merr.) Genotypes for Drought Tolerance and Genetic Diversity Analysis Using SSR Markers. Scientist 2023, 3, 31–46. [Google Scholar]
  11. Tripathi, N.; Tripathi, M.K.; Tiwari, S.; Payasi, D.K. Molecular Breeding to Overcome Biotic Stresses in Soybean: Update. Plants 2022, 11, 1967. [Google Scholar] [CrossRef] [PubMed]
  12. Yadav, R.K.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Asati, R.; Patel, V.; Sikarwar, R.S.; Payasi, D.K. Breeding and Genomic Approaches Towards Development of Fusarium Wilt Resistance in Chickpea. Life 2023, 13, 988. [Google Scholar] [CrossRef] [PubMed]
  13. Asati, R.; Tripathi, M.K.; Tiwari, S.; Yadav, R.K.; Tripathi, N. Molecular Breeding and Drought Tolerance in Chickpea. Life 2022, 12, 1846. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, H.; Zhang, J.; Lang, Z.; Botella, J.R.; Zhu, J.-K. Genome Editing—Principles and Applications for Functional Genomics Research and Crop Improvement. Crit. Rev. Plant Sci. 2017, 36, 291–309. [Google Scholar] [CrossRef]
  15. Mishra, R.; Zhao, K. Genome Editing Technologies and Their Applications in Crop Improvement. Plant Biotechnol. Rep. 2018, 12, 57–68. [Google Scholar] [CrossRef]
  16. Imran, M.; Butt, M.; Hannan, A.; Manzoor, A.; Qaisar, U. Gene Editing: A Potential Tool to Enhance Field Crop Production. Int. J. Biotech Trends Technol. 2020, 10, 72–82. [Google Scholar] [CrossRef]
  17. Tan, Y.Y.; Du, H.; Wu, X.; Liu, Y.H.; Jiang, M.; Song, S.Y.; Wu, L.; Shu, Q.Y. Gene Editing: An Instrument for Practical Application of Gene Biology to Plant Breeding. J. Zhejiang Univ. Sci. B 2020, 21, 460–473. [Google Scholar] [CrossRef] [PubMed]
  18. Abdurakhmonov, I.Y.; Ayubov, M.S.; Ubaydullaeva, K.A.; Buriev, Z.T.; Shermatov, S.E.; Ruziboev, H.S.; Shapulatov, U.M.; Saha, S.; Ulloa, M.; Yu, J.Z.; et al. RNA interference for Functional Genomics and Improvement of Cotton (Gossypium spp.). Front. Plant Sci. 2016, 7, 202. [Google Scholar] [CrossRef] [Green Version]
  19. Mohanta, T.K.; Bashir, T.; Hashem, A.; Abd_Allah, E.F.; Bae, H. Genome editing tools in plants. Genes 2017, 8, 399. [Google Scholar] [CrossRef] [Green Version]
  20. Gaj, T.; Gersbach, C.A.; Barbas, C.F., III. ZFN, TALEN and CRISPR/Cas-Based Methods for Genome Engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, K.; Gao, C. Targeted Genome Modification Technologies and Their Applications in Crop Improvements. Plant Cell Rep. 2014, 33, 575–583. [Google Scholar] [CrossRef]
  22. 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] [Green Version]
  23. Rathore, M.S.; Tiwari, S.; Tripathi, N.; Tripathi, M.K.; Tiwari, S. Status and Scenario of Genome Editing Device CRISPR-Cas9 in Crop Advancement. Curr. Appl. Sci. Technol. 2021, 40, 8–20. [Google Scholar] [CrossRef]
  24. Tripathi, L.; Dhugga, K.S.; Ntui, V.O.; Runo, S.; Syombua, E.D.; Muiruri, S.; Wen, Z.; Tripathi, J.N. Genome Editing for Sustainable Agriculture in Africa. Front. Genome Ed. 2022, 4, 876697. [Google Scholar] [CrossRef]
  25. Nadakuduti, S.S.; Buell, C.R.; Voytas, D.F.; Starker, C.G.; Douches, D.S. Genome Editing for Crop Improvement—Applications in Clonally Propagated Polyploids with a Focus on Potato (Solanum tuberosum L.). Front. Plant Sci. 2018, 9, 1607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Pixley, K.V.; Falck-Zepeda, J.B.; Paarlberg, R.L.; Phillips, P.W.B.; Slamet-Loedin, I.H.; Dhugga, K.S.; Campos, H.; Gutterson, N. Genome-Edited Crops for Improved Food Security of Smallholder Farmers. Nat. Genet. 2022, 54, 364–367. [Google Scholar] [CrossRef] [PubMed]
  27. Capecchi, M.R. High Efficiency Transformation by Direct Microinjection of DNA into Cultured Mammalian Cells. Cell 1980, 22 Pt 2, 479–488. [Google Scholar] [CrossRef]
  28. Bhattacharya, A.; Parkhi, V.; Char, B. Genome Editing for Crop Improvement: A Perspective from India. In Vitro Cell. Dev. Biol. Plant 2021, 57, 565–573. [Google Scholar] [CrossRef]
  29. Ferreira, S.S.; Reis, R.S. Using CRISPR/Cas to Enhance Gene Expression for Crop Trait Improvement by Editing miRNA Targets. J. Exp. Bot. 2023, 74, 2208–2212. [Google Scholar] [CrossRef]
  30. Jansing, J.; Schiermeyer, A.; Schillberg, S.; Fischer, R.; Bortesi, L. Genome Editing in Agriculture: Technical and Practical Considerations. Int. J. Mol. Sci. 2019, 20, 2888. [Google Scholar] [CrossRef] [Green Version]
  31. El-Mounadi, K.; Morales-Floriano, M.L.; Garcia-Ruiz, H. Principles, Applications, and Biosafety of Plant Genome Editing Using CRISPR-Cas9. Front. Plant Sci. 2020, 11, 56. [Google Scholar] [CrossRef]
  32. Upadhyay, S.; Singh, A.K.; Tripathi, M.K.; Tiwari, S.; Tripathi, N. Biotechnological Interventions to Combat Against Charcoal Rot and Rhizoctonia Root Rot Diseases of Soybean [Glycine max (L.) Merrill]. Curr. Top. Agric. Sci. 2022, 6, 1–18. [Google Scholar] [CrossRef]
  33. Rajpoot, P.; Tripathi, M.K.; Tiwari, S.; Bimal, S.S.; Tripathi, N.; Parihar, P.; Pandya, R.K.; Satyavathi, C.T. Characterization of Pearl Millet [Pennisetum glaucum (L.) R Br.] Genotypes Against Blast Disease Employing Disease Scoring and Gene Specific SSR Markers. Scientist 2023, 3, 16–30. [Google Scholar]
  34. Mishra, N.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Gupta, N.; Sharma, A.; Solanki, R.S.; Tiwari, S. Characterization of Soybean Genotypes on the Basis of Yield Attributing Traits and SSR Molecular Markers. In Innovations in Science and Technology; B P International: Hong Kong, China, 2022; Volume 3, pp. 87–106. [Google Scholar] [CrossRef]
  35. Yadav, P.K.; Singh, A.K.; Tripathi, M.K.; Tiwari, S.; Yadav, S.K.; Tripathi, N. Morphophysiological and Molecular Characterization of Maize (Zea mays L.) Genotypes for Drought Tolerance. Eur. J. Appl. Sci. 2022, 10, 65–87. [Google Scholar]
  36. Hoang, T.M.L.; Tran, T.N.; Nguyen, T.K.T.; Williams, B.; Wurm, P.; Bellairs, S.; Mundree, S. Improvement of Salinity Stress Tolerance in Rice: Challenges and Opportunities. Agronomy 2016, 6, 54. [Google Scholar] [CrossRef] [Green Version]
  37. Zaidi, S.S.-e.-A.; Mahas, A.; Vanderschuren, H.; Mahfouz, M.M. Engineering crops of the future: CRISPR approaches to develop climate-resilient and disease-resistant plants. Genome Biol. 2020, 21, 289. [Google Scholar] [CrossRef] [PubMed]
  38. Schaart, J.G.; van de Wiel, C.C.M.; Lotz, L.A.P.; Smulders, M.J.M. Opportunities for Products of New Plant Breeding Techniques. Trends Plant Sci. 2016, 21, 438–449. [Google Scholar] [CrossRef]
  39. Hazman, M.Y. Are CRISPR/Cas genome editing techniques the future of plant breeding? Egypt. J. Agril. Res. 2022, 101, 1–13. [Google Scholar] [CrossRef]
  40. Nekrasov, V.; Staskawicz, B.; Weigel, D.; Jones, J.D.; Kamoun, S. Targeted Mutagenesis in the Model Plant Nicotiana benthamiana Using Cas9 RNA-Guided Endonuclease. Nat. Biotechnol. 2013, 31, 691–693. [Google Scholar] [CrossRef]
  41. Li, J.F.; Norville, J.E.; Aach, J.; McCormack, M.; Zhang, D.; Bush, J.; Church, G.M.; Sheen, J. Multiplex and Homologous Recombination-Mediated Genome Editing in Arabidopsis and Nicotiana benthamiana Using Guide RNA and Cas9. Nat. Biotechnol. 2013, 31, 688–691. [Google Scholar] [CrossRef]
  42. 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] [PubMed]
  43. Chilcoat, D.; Liu, Z.B.; Sander, J. Use of CRISPR/Cas9 for Crop Improvement in Maize and Soybean. Prog Mol Biol Transl Sci. 2017, 149, 27–46. [Google Scholar] [CrossRef] [PubMed]
  44. Li, H.; Yang, Y.; Hong, W.; Huang, M.; Wu, M.; Zhao, X. Applications of Genome Editing Technology in the Targeted Therapy of Human Diseases: Mechanisms, Advances and Prospects. Signal Transduct. Target. Ther. 2020, 5, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhang, Y.; Showalter, A.M. CRISPR/Cas9 genome editing technology: A valuable tool for understanding plant cell wall biosynthesis and function. Front. Plant Sci. 2020, 11, 589517. [Google Scholar] [CrossRef]
  46. Gao, C. Genome Engineering for Crop Improvement and Future Agriculture. Cell 2021, 184, 1621–1635. [Google Scholar] [CrossRef]
  47. Carroll, D. Genome Engineering with Zinc-Finger Nucleases. Genetics 2011, 188, 773–782. [Google Scholar] [CrossRef] [Green Version]
  48. Urnov, F.D.; Rebar, E.J.; Holmes, M.C.; Zhang, H.S.; Gregory, P.D. Genome Editing with Engineered Zinc Finger Nucleases. Nat. Rev. Genet. 2010, 11, 636–646. [Google Scholar] [CrossRef]
  49. Bibikova, M.; Carroll, D.; Segal, D.J.; Trautman, J.K.; Smith, J.; Kim, Y.G.; Chandrasegaran, S. Stimulation of Homologous Recombination Through Targeted Cleavage by Chimeric Nucleases. Mol. Cell. Biol. 2001, 21, 289–297. [Google Scholar] [CrossRef] [Green Version]
  50. Puchta, H. The Repair of Double-Strand Breaks in Plants: Mechanisms and Consequences for Genome Evolution. J. Exp. Bot. 2005, 56, 1–14. [Google Scholar] [CrossRef] [Green Version]
  51. Lusser, M.; Parisi, C.; Plan, D.; Rodriguez-Cerezo, E. New Plant Breeding Techniques. In En State-of-the-Art and Prospects for Commercial Development 1-220; Joint Research Centre-Institute for Prospective Technological Studies, Publications Office of the European Union: Luxembourg, 2011; p. 24760. [Google Scholar]
  52. Araki, M.; Nojima, K.; Ishii, T. Caution Required for Handling Genome Editing Technology. Trends Biotechnol. 2014, 32, 234–237. [Google Scholar] [CrossRef] [Green Version]
  53. 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]
  54. Townsend, J.A.; Wright, D.A.; Winfrey, R.J.; Fu, F.; Maeder, M.L.; Joung, J.K.; Voytas, D.F. High-Frequency Modification of Plant Genes Using Engineered Zinc-Finger Nucleases. Nature 2009, 459, 442–445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zhang, F.; Voytas, D.F. Targeted Mutagenesis in Arabidopsis Using Zinc-Finger Nucleases. In Plant Chromosome Engineering; Humana Press: Totowa, NJ, USA, 2011; Volume 701, pp. 167–177. [Google Scholar] [CrossRef]
  56. Curtin, S.J.; Anderson, J.E.; Starker, C.G.; Baltes, N.J.; Mani, D.; Voytas, D.F.; Stupar, R.M. Targeted Mutagenesis for Function Analysis of Gene Duplication in Legumes. In Legume Genomics; Humana Press: Totowa, NJ, USA, 2013; pp. 25–42. [Google Scholar]
  57. Petolino, J.F. Genome Editing in Plants via Designed Zinc Finger Nucleases. In Vitro Cell. Dev. Biol. Plant 2015, 51, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Khalil, A.M. The genome editing revolution: Review. J. Gernt. Engin. Biotechnol. 2020, 18, 68. [Google Scholar] [CrossRef]
  59. Arnould, S.; Delenda, C.; Grizot, S.; Desseaux, C.; Pâques, F.; Silva, G.H.; Smith, J. The I-CreI Meganuclease and Its Engineered Derivatives: Applications from Cell Modification to Gene Therapy. Protein Eng. Des. Sel. 2011, 24, 27–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Silva, G.; Poirot, L.; Galetto, R.; Smith, J.; Montoya, G.; Duchateau, P.; Pâques, F. Mega Nucleases and Other Tools for Targeted Genome Engineering: Perspectives and Challenges for Gene Therapy. Curr. Gene Ther. 2011, 11, 11–27. [Google Scholar] [CrossRef] [Green Version]
  61. Belfort, M.; Bonocora, R.P. Homing Endonucleases: From Genetic Anomalies to Programmable Genomic Clippers. Methods Mol. Biol. 2014, 1123, 1–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sultan, L.D.; Mileshina, D.; Grewe, F.; Rolle, K.; Abudraham, S.; Głodowicz, P.; Niazi, A.K.; Keren, I.; Shevtsov, S.; Klipcan, L.; et al. The Reverse Transcriptase/RNA Maturase Protein MatR Is Required for the Splicing of Various Group II Introns in Brassicaceae Mitochondria. Plant Cell. 2016, 28, 2805–2829. [Google Scholar] [CrossRef] [Green Version]
  63. Kaur, N.; Sharma, S.; Hasanuzzaman, M.; Pati, P.K. Genome Editing: A Promising Approach for Achieving Abiotic Stress Tolerance in Plants. Int. J. Genom. 2022, 2022, 5547231. [Google Scholar] [CrossRef]
  64. Joung, J.K.; Sander, J.D. TALENs: A Widely Applicable Technology for Targeted Genome Editing. Nat. Rev. Mol. Cell Biol. 2013, 14, 49–55. [Google Scholar] [CrossRef] [Green Version]
  65. Martínez-Fortún, 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] [PubMed]
  66. Malzahn, A.; Lowder, L.; Qi, Y. Plant Genome Editing with TALEN and CRISPR. Cell Biosci. 2017, 7, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. 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]
  68. Haun, W.; Coffman, A.; Clasen, B.M.; Demorest, Z.L.; Lowy, A.; Ray, E.; Retterath, A.; Stoddard, T.; Juillerat, A.; Cedrone, F.; et al. Improved Soybean Oil Quality by Targeted Mutagenesis of the Fatty Acid Desaturase 2 Gene Family. Plant Biotechnol. J. 2014, 12, 934–940. [Google Scholar] [CrossRef]
  69. 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] [Green Version]
  70. Boch, J.; Bonas, U. Xanthomonas AvrBs3 Family-Type III Effectors: Discovery and Function. Annu. Rev. Phytopathol. 2010, 48, 419–436. [Google Scholar] [CrossRef]
  71. Bogdanove, A.J.; Schornack, S.; Lahaye, T. TAL Effectors: Finding Plant Genes for Disease and Defense. Curr. Opin. Plant Biol. 2010, 13, 394–401. [Google Scholar] [CrossRef]
  72. 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]
  73. 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]
  74. Zhou, J.; Li, D.; Wang, G.; Wang, F.; Kunjal, M.; Joldersma, D.; Liu, Z. Application and Future Perspective of CRISPR/Cas9 Genome Editing in Fruit Crops. J. Integr. Plant Biol. 2020, 62, 269–286. [Google Scholar] [CrossRef] [Green Version]
  75. Ma, X.; Liu, Y.G. CRISPR/Cas9-Based Multiplex Genome Editing in Monocot and Dicot Plants. Curr. Protoc. Mol. Biol. 2016, 115, 31.6.1–31.6.21. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, D.; Li, Z.; Li, J.F. Targeted Gene Manipulation in Plants Using the CRISPR/Cas Technology. J. Genet. Genom. 2016, 43, 251–262. [Google Scholar] [CrossRef]
  77. Ricroch, A.; Clairand, P.; Harwood, W. Use of CRISPR Systems in Plant Genome Editing: Toward New Opportunities in Agriculture. Emerg. Top. Life Sci. 2017, 1, 169–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Makarova, K.S.; Haft, D.H.; Barrangou, R.; Brouns, S.J.; Charpentier, E.; Horvath, P.; Moineau, S.; Mojica, F.J.; 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] [PubMed] [Green Version]
  79. Zhang, H.X.; Zhang, Y.; Yin, H. Genome Editing with mRNA Encoding ZFN, TALEN, and Cas9. Mol. Ther. 2019, 27, 735–746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Bortesi, L.; Fischer, R. The CRISPR/Cas9 System for Plant Genome Editing and Beyond. Biotechnol. Adv. 2015, 33, 41–52. [Google Scholar] [CrossRef]
  81. 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]
  82. Ming, M.; Ren, Q.; Pan, C.; He, Y.; Zhang, Y.; Liu, S.; Zhong, Z.; Wang, J.; Malzahn, A.A.; Wu, J.; et al. CRISPR-Cas12benables Efficient Plant Genome Engineering. Nat. Plants 2020, 6, 202–208. [Google Scholar] [CrossRef]
  83. Cong, L.; Zhang, F. Genome Engineering Using CRISPR-Cas9 System. Methods Mol. Biol. 2015, 1239, 197–217. [Google Scholar] [CrossRef] [Green Version]
  84. Tan, J.; Zhao, Y.; Wang, B.; Hao, Y.; Wang, Y.; Li, Y.; Luo, W.; Zong, W.; Li, G.; Chen, S.; et al. Efficient CRISPR/Cas9-Based Plant Genomic Fragment Deletions by Microhomology-Mediated End Joining. Plant Biotechnol. J. 2020, 18, 2161–2163. [Google Scholar] [CrossRef] [Green Version]
  85. Tang, X.; Lowder, L.G.; Zhang, T.; Malzahn, A.A.; Zheng, X.; Voytas, D.F.; Zhong, Z.; Chen, Y.; Ren, Q.; Li, Q.; et al. A CRISPR-Cpf1 System for Efficient Genome Editing and Transcriptional Repression in Plants. Nat. Plants 2017, 3, 17103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wang, M.; Xu, Z.; Gosavi, G.; Ren, B.; Cao, Y.; Kuang, Y.; Zhou, C.; Spetz, C.; Yan, F.; Zhou, X.; et al. Targeted Base Editing in Rice with CRISPR/ScCas9 System. Plant Biotechnol. J. 2020, 18, 1645–1647. [Google Scholar] [CrossRef] [PubMed]
  87. 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. [Google Scholar] [CrossRef]
  88. Shan, Q.; Wang, Y.; Li, J.; Gao, C. Genome Editing in Rice and Wheat Using the CRISPR/Cas System. Nat. Protoc. 2014, 9, 2395–2410. [Google Scholar] [CrossRef] [PubMed]
  89. Kim, D.; Alptekin, B.; Budak, H. CRISPR/Cas9 Genome Editing in Wheat. Funct. Integr. Genom. 2018, 18, 31–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Guha, T.K.; Edgell, D.R. Applications of Alternative Nucleases in the Age of CRISPR/Cas9. Int. J. Mol. Sci. 2017, 18, 2565. [Google Scholar] [CrossRef] [Green Version]
  91. Sauer, N.J.; Mozoruk, J.; Miller, R.B.; Warburg, Z.J.; Walker, K.A.; Beetham, P.R.; Schöpke, C.R.; Gocal, G.F. Oligonucleotide Directed Mutagenesis for Precision Gene Editing. Plant Biotechnol. J. 2016, 14, 496–502. [Google Scholar] [CrossRef] [Green Version]
  92. Yang, S.H.; Kim, E.; Park, H.; Koo, Y. Selection of the High Efficient sgRNA for CRISPR-Cas9 to Edit Herbicide Related Genes, PDS, ALS, and EPSPS in Tomato. Appl. Biol. Chem. 2022, 65, 13. [Google Scholar] [CrossRef]
  93. Janik, E.; Niemcewicz, M.; Ceremuga, M.; Krzowski, L.; Saluk-Bijak, J.; Bijak, M. Various Aspects of a Gene Editing System crispr–cas 9. Int. J. Mol. Sci. 2020, 21, 9604. [Google Scholar] [CrossRef]
  94. Eş, I.; Gavahian, M.; Marti-Quijal, F.J.; Lorenzo, J.M.; Mousavi Khaneghah, A.M.; Tsatsanis, C.; Kampranis, S.C.; Barba, F.J. The Application of the CRISPR-Cas9 Genome Editing Machinery in Food and Agricultural Science: Current Status, Future Perspectives, and Associated Challenges. Biotechnol. Adv. 2019, 37, 410–421. [Google Scholar] [CrossRef]
  95. Noman, A.; Aqeel, M.; He, S. CRISPR-Cas9: Tool for Qualitative and Quantitative Plant Genome Editing. Front. Plant Sci. 2016, 7, 1740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Mao, Y.; Zhang, H.; Xu, N.; Zhang, B.; Gou, F.; Zhu, J.K. Application of the CRISPR-Cas System for Efficient Genome Engineering in Plants. Mol. Plant 2013, 6, 2008–2011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Eid, A.; Mahfouz, M.M. Genome Editing: The Road of CRISPR/Cas9 from Bench to Clinic. Exp. Mol. Med. 2016, 48, e265. [Google Scholar] [CrossRef] [Green Version]
  98. Weeks, D.P.; Spalding, M.H.; Yang, B. Use of Designer Nucleases for Targeted Gene and Genome Editing in Plants. Plant Biotechnol. J. 2016, 14, 483–495. [Google Scholar] [CrossRef]
  99. Cho, S.W.; Kim, S.; Kim, J.M.; Kim, J.S. Targeted Genome Engineering in Human Cells with the Cas9 RNA-Guided Endonuclease. Nat. Biotechnol. 2013, 31, 230–232. [Google Scholar] [CrossRef] [PubMed]
  100. Prajapat, R.K.; Mathur, M.; Upadhyay, T.K.; Lal, D.; Maloo, S.; Sharma, D. Genome Editing for Crop Improvement. In Crop Improvement; CRC Press: Boca Raton, FL, USA, 2021; pp. 111–123. [Google Scholar]
  101. Khandagale, K.; Nadaf, A. Genome Editing for Targeted Improvement of Plants. Plant Biotechnol. Rep. 2016, 10, 327–343. [Google Scholar] [CrossRef]
  102. Cantos, C.; Francisco, P.; Trijatmiko, K.R.; Slamet-Loedin, I.; Chadha-Mohanty, P.K. Identification of “Safe Harbor” Loci in Indica Rice Genome by Harnessing the Property of Zinc-Finger Nucleases to Induce DNA Damage and Repair. Front. Plant Sci. 2014, 5, 302. [Google Scholar] [CrossRef] [Green Version]
  103. Shan, Q.; Wang, Y.; Chen, K.; Liang, Z.; Li, J.; Zhang, Y.; Zhang, K.; Liu, J.; Voytas, D.F.; Zheng, X.; et al. Rapid and Efficient Gene Modification in Rice and Brachypodium Using TALENs. Mol. Plant. 2013, 6, 1365–1368. [Google Scholar] [CrossRef] [Green Version]
  104. Zhang, H.; Gou, F.; Zhang, J.; Liu, W.; Li, Q.; Mao, Y.; Botella, J.R.; Zhu, J.K. TALEN-Mediated Targeted Mutagenesis Produces a Large Variety of Heritable Mutations in Rice. Plant Biotechnol. J. 2016, 14, 186–194. [Google Scholar] [CrossRef]
  105. 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]
  106. Schornack, S.; Meyer, A.; Römer, P.; Jordan, T.; Lahaye, T. Gene-for-Gene-Mediated Recognition of Nuclear-Targeted AvrBs3-Like Bacterial Effector Proteins. J. Plant Physiol. 2006, 163, 256–272. [Google Scholar] [CrossRef]
  107. Ainley, W.M.; Sastry-Dent, L.; Welter, M.E.; Murray, M.G.; Zeitler, B.; Amora, R.; Corbin, D.R.; Miles, R.R.; Arnold, N.L.; Strange, T.L.; et al. Trait Stacking via Targeted Genome Editing. Plant Biotechnol. J. 2013, 11, 1126–1134. [Google Scholar] [CrossRef]
  108. Liang, Z.; Zhang, K.; Chen, K.; Gao, C. Targeted Mutagenesis in Zea mays Using TALENs and the CRISPR/Cas System. J. Genet. Genom. 2014, 41, 63–68. [Google Scholar] [CrossRef] [PubMed]
  109. Djukanovic, V.; Smith, J.; Lowe, K.; Yang, M.; Gao, H.; Jones, S.; Nicholson, M.G.; West, A.; Lape, J.; Bidney, D.; et al. Male-Sterile Maize Plants Produced by Targeted Mutagenesis of the Cytochrome P450-Like Gene (MS26) Using a Re-designed I-CreI Homing Endonuclease. Plant J. 2013, 76, 888–899. [Google Scholar] [CrossRef]
  110. Wendt, T.; Holm, P.B.; Starker, C.G.; Christian, M.; Voytas, D.F.; Brinch-Pedersen, H.; Holme, I.B. TAL Effector Nucleases Induce Mutations at a Pre-selected Location in the Genome of Primary Barley Transformants. Plant Mol. Biol. 2013, 83, 279–285. [Google Scholar] [CrossRef]
  111. Curtin, S.J.; Zhang, F.; Sander, J.D.; Haun, W.J.; Starker, C.; Baltes, N.J.; Reyon, D.; Dahlborg, E.J.; Goodwin, M.J.; Coffman, A.P.; et al. Targeted Mutagenesis of Duplicated Genes in Soybean with Zinc-Finger Nucleases. Plant Physiol. 2011, 156, 466–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Wright, D.A.; Townsend, J.A.; Winfrey, R.J., Jr.; Irwin, P.A.; Rajagopal, J.; Lonosky, P.M.; Hall, B.D.; Jondle, M.D.; Voytas, D.F. High-Frequency Homologous Recombination in Plants Mediated by Zinc-Finger Nucleases. Plant J. 2005, 44, 693–705. [Google Scholar] [CrossRef] [PubMed]
  113. Novak, S. Plant Biotechnology Applications of Zinc Finger Technology. In Transgenic Plants; Humana Press: Totowa, NJ, USA, 2019; pp. 295–310. [Google Scholar] [CrossRef]
  114. Gago, C.; Drosou, V.; Paschalidis, K.; Guerreiro, A.; Miguel, G.; Antunes, D.; Hilioti, Z. Targeted Gene Disruption Coupled with Metabolic Screen Approach to Uncover the LEAFY COTYLEDON1-LIKE4 (L1L4) Function in Tomato Fruit Metabolism. Plant Cell Rep. 2017, 36, 1065–1082. [Google Scholar] [CrossRef]
  115. 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] [Green Version]
  116. Clasen, B.M.; Stoddard, T.J.; Luo, S.; Demorest, Z.L.; Li, J.; Cedrone, F.; Tibebu, R.; Davison, S.; Ray, E.E.; Daulhac, A.; et al. Improving Cold Storage and Processing Traits in Potato Through Targeted Gene Knockout. Plant Biotechnol. J. 2016, 14, 169–176. [Google Scholar] [CrossRef]
  117. Geurts, M.H.; Clevers, H. CRISPR Engineering in Organoids for Gene Repair and Disease Modelling. Nat. Rev. Bioeng. 2023, 1, 32–45. [Google Scholar] [CrossRef]
  118. Komor, A.C.; Zhao, K.T.; Packer, M.S.; Gaudelli, N.M.; Waterbury, A.L.; Koblan, L.W.; Kim, Y.B.; Badran, A.H.; Liu, D.R. Improved Base Excision Repair Inhibition and Bacteriophage Mu Gam Protein Yields C:G-to-T:A Base Editors with Higher Eficiency and Product Purity. Sci. Adv. 2017, 3, eaao4774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Cascalho, M. Advantages and disadvantages of cytidine deamination. J. Immunol. 2004, 172, 6513–6518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Levy, J.M.; Yeh, W.H.; Pendse, N.; Davis, J.R.; Hennessey, E.; Butcher, R.; Koblan, L.W.; Comander, J.; Liu, Q.; Liu, D.R. Cytosine and Adenine Base Editing of the Brain, Liver, Retina, Heart and Skeletal Muscle of Mice via Adeno-Associated Viruses. Nat. Biomed. Eng. 2020, 4, 97–110. [Google Scholar] [CrossRef]
  121. Anzalone, A.V.; Gao, X.D.; Podracky, C.J.; Nelson, A.T.; Koblan, L.W.; Raguram, A.; Levy, J.M.; Mercer, J.A.M.; Liu, D.R. Programmable deletion, replacement, integration and inversion of large DNA sequences with twin prime editing. Nat. Biotechnol. 2022, 40, 731–740. [Google Scholar] [CrossRef]
  122. Nelson, J.W.; Randolph, P.B.; Shen, S.P.; Everette, K.A.; Chen, P.J.; Anzalone, A.V.; An, M.; Newby, G.A.; Chen, J.C.; Hsu, A.; et al. Engineered pegRNAs Improve Prime Editing Eficiency. Nat. Biotechnol. 2022, 40, 402–410. [Google Scholar] [CrossRef]
  123. Chen, P.J.; Hussmann, J.A.; Yan, J.; Knipping, F.; Ravisankar, P.; Chen, P.F.; Chen, C.; Nelson, J.W.; Newby, G.A.; Sahin, M.; et al. Enhanced Prime Editing Systems by Manipulating Cellular Determinants of Editing Outcomes. Cell 2021, 184, 5635–5652. [Google Scholar] [CrossRef]
  124. Liu, Z.; Ma, C.; Hou, L.; Wu, X.; Wang, D.; Zhang, L.; Liu, P. Exogenous SA Affects Rice Seed Germination Under Salt Stress by Regulating Na(+)/K(+) Balance and Endogenous GAs and ABA Homeostasis. Int. J. Mol. Sci. 2022, 23, 3293. [Google Scholar] [CrossRef]
  125. Ningwal, R.; Tripathi, M.K.; Tiwari, S.; Yadav, R.K.; Tripathi, N.; Solanki, R.S.; Asati, R.; Yasin, M. Assessment of Genetic Variability, Correlation and Path Coefficient Analysis for Yield and Its Attributing Traits in Chickpea (Cicer arietinum L.) The Pharma. Innov. J. 2023, 12, 4851–4859. [Google Scholar]
  126. Sharma, S.; Tripathi, M.K.; Tiwari, S.; Solanki, R.S.; Chauhan, S.; Tripathi, N.; Dwivedi, N.; Kandalkar, V.S. The Exploitation of Genetic Variability and Trait Association Analysis for Diverse Quantitative Traits in Bread Wheat (Triticum aestivum L.). Curr. J. Appl. Sci. Technol. 2023, 42, 19–33. [Google Scholar] [CrossRef]
  127. Shyam, C.; Tripathi, M.K.; Tripathi, N.; Tiwari, S.; Sikarwar, R.S. Identification of Low and High Erucic Acid Containing Genotype(S) in Indian Mustard Employing Molecular Markers. In Recent Progress in Plant and Soil Research; B P International: Hong Kong, China, 2022; Volume 5, pp. 18–36. [Google Scholar] [CrossRef]
  128. Yadav, S.; Tiwari, S.; Tripathi, M.K.; Tripathi, N.; Gupta, N.; Tiwari, S. Evaluation of High Oleic Acid Content in a Set of 96 Genotypes of Arachis hypogaea L. Scientist 2023, 2, 132–143. [Google Scholar]
  129. Sandhu, K.S.; Mihalyov, P.D.; Lewien, M.J.; Pumphrey, M.O.; Carter, A.H. Genomic Selection and Genome-Wide Association Studies for Grain Protein Content Stability in a Nested Association Mapping Population of Wheat. Agronomy 2021, 11, 2528. [Google Scholar] [CrossRef]
  130. Sandhu, K.S.; Patil, S.S.; Aoun, M.; Carter, A.H. Multi-trait Multi-environment Genomic Prediction for End-Use Quality Traits in Winter Wheat. Front. Genet. 2022, 13, 831020. [Google Scholar] [CrossRef]
  131. 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]
  132. He, M.; He, C.Q.; Ding, N.Z. Abiotic Stresses: General Defenses of Land Plants and Chances for Engineering Multistress Tolerance. Front. Plant Sci. 2018, 9, 1771. [Google Scholar] [CrossRef] [Green Version]
  133. Joshi, R.K.; Bharat, S.S.; Mishra, R. Engineering Drought Tolerance in Plants Through CRISPR/Cas Genome Editing. 3 Biotech 2020, 10, 400. [Google Scholar] [CrossRef]
  134. Kumar, S. Abiotic Stresses and Their Effects on Plant Growth, Yield and Nutritional Quality of Agricultural Produce. Int. J. Food Sci. Agric. 2020, 4, 367–378. [Google Scholar] [CrossRef]
  135. Daryanto, S.; Wang, L.; Jacinthe, P.A. Global Synthesis of Drought Effects on Maize and Wheat Production. PLoS ONE 2016, 11, e0156362. [Google Scholar] [CrossRef] [Green Version]
  136. Daryanto, S.; Wang, L.; Jacinthe, P.-A. Global Synthesis of Drought Effects on Cereal, Legume, Tuber and Root Crops Production: A Review. Agric. Water Manag. 2017, 179, 18–33. [Google Scholar] [CrossRef] [Green Version]
  137. Mafakheri, A.; Siosemardeh, A.; Bahramnejad, B.; Struik, P.C.; Sohrabi, Y. Effect of Drought Stress on Yield, Proline and Chlorophyll Contents in Three Chickpea Cultivars. Aust. J. Crop Sci. 2010, 4, 580–585. [Google Scholar]
  138. Farooq, M.; Gogoi, N.; Barthakur, S.; Baroowa, B.; Bharadwaj, N.; Alghamdi, S.S.; Siddique, K.H.M. Drought Stress in Grain Legumes During Reproduction and Grain Filling. J. Agron. Crop Sci. 2017, 203, 81–102. [Google Scholar] [CrossRef]
  139. Maleki, A.; Naderi, A.; Naseri, R.; Fathi, A.; Bahamin, S.; Maleki, R. Physiological Performance of Soybean Cultivars Under Drought Stress. Bull. Environ. Pharmacol. Life Sci. 2013, 2, 38–44. [Google Scholar]
  140. 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]
  141. Tian, H.; Chen, S.; Yang, W.; Wang, T.; Zheng, K.; Wang, Y.; Cheng, Y.; Zhang, N.; Liu, S.; Li, D.; et al. A Novel Family of Transcription Factors Conserved in Angiosperms Is Required for ABA Signalling. Plant Cell Environ. 2017, 40, 2958–2971. [Google Scholar] [CrossRef]
  142. 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]
  143. Zhao, Y.; Zhang, C.; Liu, W.; Gao, W.; Liu, C.; Song, G.; Li, W.X.; Mao, L.; Chen, B.; Xu, Y.; et al. An Alternative Strategy for Targeted Gene Replacement in Plants Using a Dual-sgRNA/Cas9 Design. Sci. Rep. 2016, 6, 23890. [Google Scholar] [CrossRef]
  144. Park, J.J.; Dempewolf, E.; Zhang, W.; Wang, Z.Y. RNA-Guided Transcriptional Activation via CRISPR/dCas9 Mimics Overexpression Phenotypes in Arabidopsis. PLoS ONE 2017, 12, e0179410. [Google Scholar] [CrossRef] [Green Version]
  145. Roca Paixão, J.F.; Gillet, F.X.; Ribeiro, T.P.; Bournaud, C.; Lourenço-Tessutti, I.T.; Noriega, D.D.; Melo, B.P.d.; 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]
  146. Nuñez-Muñoz, L.; Vargas-Hernández, B.; Hinojosa-Moya, J.; Ruiz-Medrano, R.; Xoconostle-Cázares, B. Plant Drought Tolerance Provided Through Genome Editing of the Trehalose Gene. Plant Signal. Behav. 2021, 16, 1877005. [Google Scholar] [CrossRef]
  147. Zhang, H.; Zhang, J.; Wei, P.; Zhang, B.; Gou, F.; Feng, Z.; Mao, Y.; Yang, L.; Zhang, H.; Xu, N.; et al. The CRISPR/Cas9 System Produces Specific and Homozygous Targeted Gene Editing in Rice in One Generation. Plant Biotechnol. J. 2014, 12, 797–807. [Google Scholar] [CrossRef]
  148. 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]
  149. 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]
  150. Usman, B.; Nawaz, G.; Zhao, N.; Liao, S.; Liu, Y.; Li, R. Precise Editing of the OsPYL9 Gene by RNA-Guided Cas9 Nuclease Confers Enhanced Drought Tolerance and Grain Yield in Rice (Oryza sativa L.) by Regulating Circadian Rhythm and Abiotic Stress Responsive Proteins. Int. J. Mol. Sci. 2020, 21, 7854. [Google Scholar] [CrossRef]
  151. 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]
  152. 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]
  153. Razzaq, M.K.; Akhter, M.; Ahmad, R.M.; Cheema, K.L.; Hina, A.; Karikari, B.; Raza, G.; Xing, G.; Gai, J.; Khurshid, M. CRISPR-Cas9 Based Stress Tolerance: New Hope for Abiotic Stress Tolerance in Chickpea (Cicer arietinum). Mol. Biol. Rep. 2022, 49, 8977–8985. [Google Scholar] [CrossRef]
  154. Guo, M.; Rupe, M.A.; Wei, J.; Winkler, C.; Goncalves-Butruille, M.; Weers, B.P.; Cerwick, S.F.; Dieter, J.A.; Duncan, K.E.; Howard, R.J.; et al. Maize ARGOS1 (ZAR1) Transgenic Alleles Increase Hybrid Maize Yield. J. Exp. Bot. 2014, 65, 249–260. [Google Scholar] [CrossRef]
  155. Shi, J.; Habben, J.E.; Archibald, R.L.; Drummond, B.J.; Chamberlin, M.A.; Williams, R.W.; Lafitte, H.R.; Weers, B.P. Overexpression of ARGOS Genes Modifies Plant Sensitivity to Ethylene, Leading to Improved Drought Tolerance in Both Arabidopsis and Maize. Plant Physiol. 2015, 169, 266–282. [Google Scholar] [CrossRef]
  156. 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]
  157. Illouz-Eliaz, N.; Nissan, I.; Nir, I.; Ramon, U.; Shohat, H.; Weiss, D. Mutations in the Tomato Gibberellin Receptors Suppress Xylem Proliferation and Reduce Water Loss Under Water-Deficit Conditions. J. Exp. Bot. 2020, 71, 3603–3612. [Google Scholar] [CrossRef]
  158. 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]
  159. 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]
  160. 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] [Green Version]
  161. Abdallah, N.A.; Elsharawy, H.; Abulela, H.A.; Thilmony, R.; Abdelhadi, A.A.; Elarabi, N.I. Multiplex CRISPR/Cas9-Mediated Genome Editing to Address Drought Tolerance in Wheat. GM Crops Food 2022, 6, 1–17. [Google Scholar] [CrossRef]
  162. He, X.; Luo, X.; Wang, T.; Liu, S.; Zhang, X.; Zhu, L. GhHB12 Negatively Regulates Abiotic Stress Tolerance in Arabidopsis and Cotton. Environ. Exp. Bot. 2020, 176, 104087. [Google Scholar] [CrossRef]
  163. Wu, J.; Yan, G.; Duan, Z.; Wang, Z.; Kang, C.; Guo, L.; Liu, K.; Tu, J.; Shen, J.; Yi, B. Roles of the Brassica napus Della Protein BnaA6. RGA, in modulating drought tolerance by interacting with the ABA signalling component BnaA10.ABF2. Front. Plant Sci. 2020, 11, 577. [Google Scholar]
  164. Asseng, S.; Ewert, F.; Martre, P.; Rötter, R.P.; Lobell, D.B.; Cammarano, D.; Kimball, B.A.; Ottman, M.J.; Wall, G.W.; White, J.W.; et al. Rising Temperatures Reduce Global Wheat Production. Nat. Clim. Chang. 2015, 5, 143–147. [Google Scholar] [CrossRef]
  165. Yang, H.; Huang, T.; Ding, M.; Lu, D.; Lu, W. High Temperature During Grain Filling Impacts on Leaf Senescence in Waxy Maize. Agron. J. 2017, 109, 906–916. [Google Scholar] [CrossRef]
  166. Wang, Y.; Wang, L.; Zhou, J.; Hu, S.; Chen, H.; Xiang, J.; Zhang, Y.; Zeng, Y.; Shi, Q.; Zhu, D.; et al. Research Progress on Heat Stress of Rice at Flowering Stage. Rice Sci. 2019, 26, 1–10. [Google Scholar] [CrossRef]
  167. Ali, M.G.M.; Ahmed, M.; Ibrahim, M.M.; El Baroudy, A.A.; Ali, E.F.; Shokr, M.S.; Aldosari, A.A.; Majrashi, A.; Kheir, A.M.S. Optimizing Sowing Window, Cultivar Choice, and Plant Density to Boost Maize Yield Under RCP8. 5 Climate Scenario of CMIP5. Int. J. Biometeorol. 2022, 66, 971–985. [Google Scholar] [CrossRef]
  168. Ejaz, M.; Abbas, G.; Fatima, Z.; Iqbal, P.; Raza, M.A.; Kheir, A.M.S.; Ahmed, M.; Kakar, K.M.; Ahmad, S. Modelling Climate Uncertainty and Adaptations for Soybean-Based Cropping System. Int. J. Plant Prod. 2022, 16, 235–250. [Google Scholar] [CrossRef]
  169. Haider, S.; Raza, A.; Iqbal, J.; Shaukat, M.; Mahmood, T. Analyzing the Regulatory Role of Heat Shock Transcription Factors in Plant Heat Stress Tolerance: A Brief Appraisal. Mol. Biol. Rep. 2022, 49, 5771–5785. [Google Scholar] [CrossRef]
  170. dos Santos, T.B.; Ribas, A.F.; de Souza, S.G.H.; Budzinski, I.G.F.; Domingues, D.S. Physiological Responses to Drought, Salinity, and Heat Stress in Plants: A Review. Stresses 2022, 2, 113–135. [Google Scholar] [CrossRef]
  171. Mathivanan, S. Abiotic Stress-Induced Molecular and Physiological Changes and Adaptive Mechanisms in Plants. In Abiotic Stress in Plants; Fahad, S., Saud, S., Chen, Y., Wu, C., Wang, D., Eds.; Intech Open: Rijeka, Croatia, 2020. [Google Scholar] [CrossRef]
  172. Boyer, J.S.; Westgate, M.E. Grain Yields with Limited Water. J. Exp. Bot. 2004, 55, 2385–2394. [Google Scholar] [CrossRef]
  173. Hazel, J.R. Thermal Adaptation in Biological Membranes: Is Homeoviscous Adaptation the Explanation? Annu. Rev. Physiol. 1995, 57, 19–42. [Google Scholar] [CrossRef]
  174. Awasthi, R.; Bhandari, K.; Nayyar, H. Temperature Stress and Redox Homeostasis in Agricultural Crops. Front. Environ. Sci. 2015, 3, 11. [Google Scholar] [CrossRef] [Green Version]
  175. Parmar, N.; Singh, K.H.; Sharma, D.; Singh, L.; Kumar, P.; Nanjundan, J.; Khan, Y.J.; Chauhan, D.K.; Thakur, A.K. Genetic Engineering Strategies for Biotic and Abiotic Stress Tolerance and Quality Enhancement in Horticultural Crops: A Comprehensive Review. 3 Biotech 2017, 7, 239. [Google Scholar] [CrossRef]
  176. 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. Biol. 2016, 90, 49–62. [Google Scholar] [CrossRef] [PubMed]
  177. Nandy, S.; Pathak, B.; Zhao, S.; Srivastava, V. Heat-Shock-Inducible CRISPR/Cas9 System Generates Heritable Mutations in Rice. Plant Direct 2019, 3, e00145. [Google Scholar] [CrossRef] [Green Version]
  178. 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] [Green Version]
  179. 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 Signalling in Tomato. Plant Cell Physiol. 2018, 59, 2239–2254. [Google Scholar] [CrossRef]
  180. 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]
  181. Li, J.; Zhang, H.; Si, X.; Tian, Y.; Chen, K.; Liu, J.; Chen, H.; Gao, C. Generation of Thermosensitive Male-Sterile Maize by Targeted Knockout of the ZmTMS5 Gene. J. Genet. Genom. 2017, 44, 465–468. [Google Scholar] [CrossRef]
  182. Bertier, L.D.; Ron, M.; Huo, H.; Bradford, K.J.; Britt, A.B.; Michelmore, R.W. High-Resolution Analysis of the Efficiency, Heritability, and Editing Outcomes of CRISPR/Cas9-Induced Modifications of NCED4 in Lettuce (Lactuca sativa). G3 Genes Genomes Genet. 2018, 8, 1513–1521. [Google Scholar] [CrossRef] [Green Version]
  183. Debbarma, J.; Sarki, Y.N.; Saikia, B.; Boruah, H.P.D.; Singha, D.L.; Chikkaputtaiah, C. Ethylene Response Factor (ERF) Family Proteins in Abiotic Stresses and CRISPR–Cas9 Genome Editing of ERFs for Multiple Abiotic Stress Tolerance in Crop Plants: A Review. Mol. Biotechnol. 2019, 61, 153–172. [Google Scholar] [CrossRef]
  184. Chang, Y.; Nguyen, B.H.; Xie, Y.; Xiao, B.; Tang, N.; Zhu, W.; Mou, T.; Xiong, L. Co-overexpression of the Constitutively Active Form of OsbZIP46 and ABA-Activated Protein Kinase SAPK6 Improves Drought and Temperature Stress Resistance in Rice. Front. Plant Sci. 2017, 8, 1102. [Google Scholar] [CrossRef] [PubMed]
  185. Ding, Y.; Shi, Y.; Yang, S. Molecular Regulation of Plant Responses to Environmental Temperatures. Mol. Plant 2020, 13, 544–564. [Google Scholar] [CrossRef] [PubMed]
  186. Yadav, S.K. Cold Stress Tolerance Mechanisms in Plants. A Review. Agron. Sustain. Dev. 2010, 30, 515–527. [Google Scholar] [CrossRef] [Green Version]
  187. Muller, O.; Stewart, J.J.; Cohu, C.M.; Polutchko, S.K.; Demmig-Adams, B.; Adams, W.W., III. Leaf Architectural, Vascular and Photosynthetic Acclimation to Temperature in Two Biennials. Physiol. Plant. 2014, 152, 763–772. [Google Scholar] [CrossRef] [PubMed]
  188. Adams, W.W., III; Stewart, J.J.; Cohu, C.M.; Muller, O.; Demmig-Adams, B. Habitat Temperature and Precipitation of Arabidopsis thaliana Ecotypes Determine the Response of Foliar Vasculature, Photosynthesis, and Transpiration to Growth Temperature. Front. Plant Sci. 2016, 7, 1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Hassan, M.A.; Xiang, C.; Farooq, M.; Muhammad, N.; Yan, Z.; Hui, X.; Yuanyuan, K.; Bruno, A.K.; Lele, Z.; Jincai, L. Cold Stress in Wheat: Plant Acclimation Responses and Management Strategies. Front. Plant Sci. 2021, 12, 676884. [Google Scholar] [CrossRef] [PubMed]
  190. Eom, S.H.; Ahn, M.A.; Kim, E.; Lee, H.J.; Lee, J.H.; Wi, S.H.; Kim, S.K.; Lim, H.B.; Hyun, T.K. Plant Response to Cold Stress: Cold Stress Changes Antioxidant Metabolism in Heading Type Kimchi Cabbage (Brassica rapa L. ssp. pekinensis). Antioxidants 2022, 11, 700. [Google Scholar] [CrossRef]
  191. Karavolias, N.G.; Horner, W.; Abugu, M.N.; Evanega, S.N. Application of Gene Editing for Climate Change in Agriculture. Front. Sustain. Food Syst. 2021, 5, 685801. [Google Scholar] [CrossRef]
  192. Ahmad, M. Plant Breeding Advancements with “CRISPR-Cas” Genome Editing Technologies Will Assist Future Food Security. Front. Plant Sci. 2023, 14, 1133036. [Google Scholar] [CrossRef] [PubMed]
  193. Lv, Y.; Yang, M.; Hu, D.; Yang, Z.; Ma, S.; Li, X.; Xiong, L. The OsMYb30 Transcription Factor Suppresses Cold Tolerance by Interacting with a JAZ Protein and Suppressing β-Amylase Expression1. Plant Physiol. 2017, 173, 1475–1491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. 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]
  195. Molla, K.A.; Shih, J.; Yang, Y. Single-Nucleotide Editing for zebra3 and wsl5 Phenotypes in Rice Using CRISPR/Cas9-Mediated Adenine Base Editors. aBIOTECH 2020, 1, 106–118. [Google Scholar] [CrossRef] [Green Version]
  196. Liu, X.; Lan, J.; Huang, Y.; Cao, P.; Zhou, C.; Ren, Y.; He, N.; Liu, S.; Tian, Y.; Nguyen, T.; et al. WSL5, a Pentatricopeptide Repeat Protein, Is Essential for Chloroplast Biogenesis in Rice Under Cold Stress. J. Exp. Bot. 2018, 69, 3949–3961. [Google Scholar] [CrossRef]
  197. Shakiba, E.; Edwards, J.D.; Jodari, F.; Duke, S.E.; Baldo, A.M.; Korniliev, P.; McCouch, S.R.; Eizenga, G.C. Genetic architecture of cold tolerance in rice (Oryza sativa) determined through high resolution genome-wide analysis. PLoS ONE 2017, 12, e0172133. [Google Scholar] [CrossRef] [Green Version]
  198. 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]
  199. Cui, Y.; Jiang, N.; Xu, Z.; Xu, Q. Heterotrimeric G Protein Are Involved in the Regulation of Multiple Agronomic Traits and Stress Tolerance in Rice. BMC Plant Biol. 2020, 20, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. 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]
  201. Lu, G.; Wang, C.; Wang, G.; Mao, G.; Habben, J.E.; Chen, G.; Liu, M.; Shi, Y.; Wang, W.; Wang, X.; et al. Knockouts of Drought Sensitive Genes Improve Rice Grain Yield Under Both Drought and Well-Watered Field Conditions. Adv. Crop Sci. Technol. 2020, 8, 444. [Google Scholar]
  202. 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]
  203. Li, J.; Meng, X.; Zong, Y.; Chen, K.; Zhang, H.; Liu, J.; Li, J.; Gao, C. Gene replacements and insertions in rice by intron targeting using CRISPR-Cas9. Nat. Plants 2016, 2, 16139. [Google Scholar] [CrossRef]
  204. Li, Q.; Zhang, D.; Chen, M.; Liang, W.; Wei, J.; Qi, Y.; Yuan, Z. Development of Japonica Photo- Sensitive Genic Male Sterile Rice Lines by Editing Carbon Starved Anther Using CRISPR/Cas9. J. Genet. Genom. 2016, 43, 415–419. [Google Scholar] [CrossRef]
  205. Zhou, H.; He, M.; Li, J.; Chen, L.; Huang, Z.; Zheng, S.; Zhu, L.; Ni, E.; Jiang, D.; Zhao, B.; et al. Development of Commercial Thermo-Sensitive Genic Male Sterile Rice Accelerates Hybrid Rice Breeding Using the CRISPR/Cas9-Mediated TMS5 Editing System. Sci. Rep. 2016, 6, 37395. [Google Scholar] [CrossRef] [Green Version]
  206. Shim, J.S.; Oh, N.; Chung, P.J.; Kim, Y.S.; Choi, Y.D.; Kim, J.K. Overexpression of OsNAC14 Improves Drought Tolerance in Rice. Front. Plant Sci. 2018, 9, 310. [Google Scholar] [CrossRef] [Green Version]
  207. 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] [PubMed]
  208. Curtin, S.J.; Xiong, Y.; Michno, J.M.; Campbell, B.W.; Stec, A.O.; Čermák, T.; Starker, C.; Voytas, D.F.; Eamens, A.L.; Stupar, R.M. Crispr/cas9 and talens generate heritable mutations for genes involved in small RNA processing of Glycine max and Medicago truncatula. Plant Biotechnol. J. 2018, 16, 1125–1137. [Google Scholar] [CrossRef] [Green Version]
  209. 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] [Green Version]
  210. 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]
  211. 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]
  212. 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] [PubMed]
  213. Chen, H.J.; Su, C.T.; Lin, C.H.; Huang, G.J.; Lin, Y.H. Expression of Sweet Potato Cysteine Protease SPCP2 Altered Developmental Characteristics and Stress Responses in Transgenic Arabidopsis Plants. J. Plant Physiol. 2010, 167, 838–847. [Google Scholar] [CrossRef]
  214. Shahid, S.A.; Zaman, M.; Heng, L. Introduction to Soil Salinity, Sodicity and Diagnostics Techniques. In Guideline for Salinity Assessment, Mitigation and Adaptation Using Nuclear and Related Techniques; Springer: Berlin/Heidelberg, Germany, 2018; pp. 1–42. [Google Scholar] [CrossRef]
  215. Gong, Z.; Xiong, L.; Shi, H.; Yang, S.; Herrera-Estrella, L.R.; Xu, G.; Chao, D.-Y.; Li, J.; Wang, P.-Y.; Qin, F. Plant Abiotic Stress Response and Nutrient Use Efficiency. Sci. China Life Sci. 2020, 63, 635–674. [Google Scholar] [CrossRef]
  216. Morton, M.J.L.; Awlia, M.; Al-Tamimi, N.; Saade, S.; Pailles, Y.; Negrão, S.; Tester, M. Salt Stress Under the Scalpel–Dissecting the Genetics of Salt Tolerance. Plant J. 2019, 97, 148–163. [Google Scholar] [CrossRef] [Green Version]
  217. Van Zelm, E.; Zhang, Y.; Testerink, C. Salt Tolerance Mechanisms of Plants. Annu. Rev. Plant Biol. 2020, 71, 403–433. [Google Scholar] [CrossRef] [Green Version]
  218. Jamil, A.; Riaz, S.; Ashraf, M.; Foolad, M.R. Gene Expression Profiling of Plants Under Salt Stress. Crit. Rev. Plant Sci. 2011, 30, 435–458. [Google Scholar] [CrossRef]
  219. Shrivastava, P.; Kumar, R. Soil Salinity: A Serious Environmental Issue and Plant Growth Promoting Bacteria as One of the Tools for Its Alleviation. Saudi J. Biol. Sci. 2015, 22, 123–131. [Google Scholar] [CrossRef] [Green Version]
  220. Gharsallah, C.; Fakhfakh, H.; Grubb, D.; Gorsane, F. Effect of Salt Stress on Ion Concentration, Proline Content, Antioxidant Enzyme Activities and Gene Expression in Tomato Cultivars. AoB Plants 2016, 8, plw055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Rahneshan, Z.; Nasibi, F.; Moghadam, A.A. Effects of Salinity Stress on Some Growth, Physiological, Biochemical Parameters and Nutrients in Two Pistachio (Pistacia vera L.) Rootstocks. J. Plant Interact. 2018, 13, 73–82. [Google Scholar] [CrossRef] [Green Version]
  222. Ali, Q.; Daud, M.K.; Haider, M.Z.; Ali, S.; Rizwan, M.; Aslam, N.; Noman, A.; Iqbal, N.; Shahzad, F.; Deeba, F.; et al. Seed Priming by Sodium Nitroprusside Improves Salt Tolerance in Wheat (Triticum aestivum L.) by Enhancing Physiological and Biochemical Parameters. Plant Physiol. Biochem. 2017, 119, 50–58. [Google Scholar] [CrossRef] [PubMed]
  223. Sahin, U.; Ekinci, M.; Ors, S.; Turan, M.; Yildiz, S.; Yildirim, E. Effects of Individual and Combined Effects of Salinity and Drought on Physiological, Nutritional and Biochemical Properties of Cabbage (Brassica oleracea var. capitata). Sci. Hortic. 2018, 240, 196–204. [Google Scholar] [CrossRef]
  224. Ghazali, E.G.E. SuaedavermiculataForssk. Ex JF Gmel: Structural Characteristics and Adaptations to Salinity and Drought: A Review. Int. J. Sci. 2020, 9, 28–33. [Google Scholar]
  225. Pan, T.; Liu, M.; Kreslavski, V.D.; Zharmukhamedov, S.K.; Nie, C.; Yu, M.; Kuznetsov, V.V.; Allakhverdiev, S.I.; Shabala, S. Non-stomatal Limitation of Photosynthesis by Soil Salinity. Crit. Rev. Environ. Sci. Technol. 2021, 51, 791–825. [Google Scholar] [CrossRef]
  226. Navada, S.; Vadstein, O.; Gaumet, F.; Tveten, A.K.; Spanu, C.; Mikkelsen, Ø.; Kolarevic, J. Biofilms Remember: Osmotic Stress Priming as a Microbial Management Strategy for Improving Salinity Acclimation in Nitrifying Biofilms. Water Res. 2020, 176, 115732. [Google Scholar] [CrossRef]
  227. Atieno, J.; Li, Y.; Langridge, P.; Dowling, K.; Brien, C.; Berger, B.; Varshney, R.K.; Sutton, T. Exploring Genetic Variation for Salinity Tolerance in Chickpea Using Image-Based Phenotyping. Sci. Rep. 2017, 7, 1300. [Google Scholar] [CrossRef] [Green Version]
  228. Huang, Y.; Guan, C.; Liu, Y.; Chen, B.; Yuan, S.; Cui, X.; Zhang, Y.; Yang, F. Enhanced Growth Performance and Salinity Tolerance in Transgenic Switchgrass via Overexpressing Vacuolar Na+ (K+)/H+ Antiporter Gene (PvNHX1). Front. Plant Sci. 2017, 8, 458. [Google Scholar] [CrossRef] [Green Version]
  229. Zhang, Q.; Dai, W. Plant Response to Salinity Stress. In Stress Physiology of Woody Plants; Dai, W., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 155–173. [Google Scholar]
  230. Munns, R.; Tester, M. Mechanisms of Salinity Tolerance. Annu. Rev. Plant Biol. 2008, 59, 651–681. [Google Scholar] [CrossRef] [Green Version]
  231. Takagi, H.; Tamiru, M.; Abe, A.; Yoshida, K.; Uemura, A.; Yaegashi, H.; Obara, T.; Oikawa, K.; Utsushi, H.; Kanzaki, E.; et al. MutMap Accelerates Breeding of a Salt-Tolerant Rice Cultivar. Nat. Biotechnol. 2015, 33, 445–449. [Google Scholar] [CrossRef]
  232. 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] [PubMed] [Green Version]
  233. Alfatih, A.; Wu, J.; Jan, S.U.; Zhang, Z.S.; Xia, J.Q.; Xiang, C.B. Loss of Rice Paraquat TOLERANCE 3 Confers Enhanced Resistance to Abiotic Stresses and Increases Grain Yield in Field. Plant Cell Environ. 2020, 43, 2743–2754. [Google Scholar] [CrossRef] [PubMed]
  234. Yue, E.; Cao, H.; Liu, B. OsmiR535, a Potential Genetic Editing Target for Drought and Salinity Stress Tolerance in Oryza sativa. Plants 2020, 9, 1337. [Google Scholar] [CrossRef]
  235. Alam, M.S.; Kong, J.; Tao, R.; Ahmed, T.; Alamin, M.; Alotaibi, S.S.; Abdelsalam, N.R.; Xu, J.H. CRISPR/Cas9 Mediated Knockout of the OsbHLH024 Transcription Factor Improves Salt Stress Resistance in Rice (Oryza sativa L.). Plants 2022, 11, 1184. [Google Scholar] [CrossRef]
  236. Liu, X.; Wu, D.; Shan, T.; Xu, S.; Qin, R.; Li, H.; Negm, M.; Wu, D.; Li, J. The Trihelix Transcription Factor OsGTc-2 Is Involved Adaption to Salt Stress in Rice. Plant Mol. Biol. 2020, 103, 545–560. [Google Scholar] [CrossRef] [PubMed]
  237. Han, X.; Chen, Z.; Li, P.; Xu, H.; Liu, K.; Zha, W.; Li, S.; Chen, J.; Yang, G.; Huang, J.; et al. Development of Novel Rice Germplasm for Salt-Tolerance at Seedling Stage Using CRISPR-Cas9. Sustainability 2022, 14, 2621. [Google Scholar] [CrossRef]
  238. 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]
  239. Wang, W.C.; Lin, T.C.; Kieber, J.; Tsai, Y.C. Response Regulators 9 and 10 Negatively Regulate Salinity Tolerance in Rice. Plant Cell Physiol. 2019, 60, 2549–2563. [Google Scholar] [CrossRef]
  240. 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]
  241. Zhang, C.; Srivastava, A.K.; Sadanandom, A. Targeted Mutagenesis of the SUMO Protease, Overly Tolerant to Salt1 in Rice Through CRISPR/Cas9-Mediated Genome Editing Reveals a Major Role of This SUMO Protease In Salt Tolerance. bioRxiv 2019. bioRxiv:555706. [Google Scholar]
  242. Lan, T.; Zheng, Y.; Su, Z.; Yu, S.; Song, H.; Zheng, X.; Lin, G.; Wu, W. OsSPL10, a SBP-Box Gene, Plays a Dual Role in Salt Tolerance and Trichome Formation in Rice (Oryza sativa L.). G3 Genes|Genomes|Genet. 2019, 9, 4107–4114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. 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]
  244. 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] [PubMed] [Green Version]
  245. 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]
  246. Vlčko, T.; Ohnoutková, L. Allelic Variants of CRISPR/Cas9 Induced Mutation in an Inositol Trisphosphate 5/6 Kinase Gene Manifest Different Phenotypes in Barley. Plants 2020, 9, E195. [Google Scholar] [CrossRef] [Green Version]
  247. 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]
  248. Guo, M.; Wang, X.S.; Guo, H.D.; Bai, S.Y.; Khan, A.; Wang, X.M.; Gao, Y.M.; Li, J.S. Tomato salt tolerance mechanisms and their potential applications for fighting salinity: A review. Front. Plant Sci. 2022, 13, 949541. [Google Scholar] [CrossRef]
  249. Alengebawy, A.; Abdelkhalek, S.T.; Qureshi, S.R.; Wang, M.Q. Heavy Metals and Pesticides Toxicity in Agricultural Soil and Plants: Ecological Risks and Human Health Implications. Toxics 2021, 9, 42. [Google Scholar] [CrossRef]
  250. Ghori, N.-H.; Ghori, T.; Hayat, M.Q.; Imadi, S.R.; Gul, A.; Altay, V.; Ozturk, M. Heavy Metal Stress and Responses in Plants. Int. J. Environ. Sci. Technol. 2019, 16, 1807–1828. [Google Scholar] [CrossRef]
  251. Aydinalp, C.; Marinova, S. The Effects of Heavy Metals on Seed Germination and Plant Growth on Alfalfa Plant (Medicago sativa). Bulg. J. Agric. Sci. 2009, 15, 347–350. [Google Scholar]
  252. Dağhan, H.; Öztürk, M.; Hakeem, K.; Sabir, M.; Mermut, A. Soil Pollution in Turkey and Remediation Methods. In Soil Remediation and Plants Prospects and Challenges; Hakeem, K.R., Sabir, M., Ozturk, M., Mermut, A.R., Eds.; Academic Press: Cambridge, MA, USA, 2015; pp. 287–312. [Google Scholar]
  253. Rascio, N.; Navari-Izzo, F. Heavy Metal Hyperaccumulating Plants: How and Why Do They Do It? and What Makes Them so Interesting? Plant Sci. 2011, 180, 169–181. [Google Scholar] [CrossRef] [PubMed]
  254. Shahid, M.; Khalid, S.; Abbas, G.; Shahid, N.; Nadeem, M.; Sabir, M.; Aslam, M.; Dumat, C. Heavy Metal Stress and Crop Productivity. In Crop Production and Global Environmental Issues; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1–25. [Google Scholar]
  255. Kaur, R.; Das, S.; Bansal, S.; Singh, G.; Sardar, S.; Dhar, H.; Ram, H. Heavy Metal Stress in Rice: Uptake, Transport, Signalling, and Tolerance Mechanisms. Physiol. Plant. 2021, 173, 430–448. [Google Scholar] [CrossRef] [PubMed]
  256. Baeg, G.J.; Kim, S.H.; Choi, D.M.; Tripathi, S.; Han, Y.J.; Kim, J. CRISPR/Cas9-Mediated Mutation of 5-Oxoprolinase Gene Confers Resistance to Sulfonamide Compounds in Arabidopsis. Plant Biotechnol. Rep. 2021, 15, 753–764. [Google Scholar] [CrossRef]
  257. Hasanuzzaman, M.; Hakeem, K.R.; Nahar, K.; Alharby, H.F. Plant Abiotic Stress Tolerance: Agronomic, Molecular and Biotechnological Approaches; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar] [CrossRef]
  258. Songmei, L.; Jie, J.; Yang, L.; Jun, M.; Shouling, X.; Yuanyuan, T.; Youfa, L.; Qingyao, S.; Jianzhong, H. Characterization and Evaluation of OsLCT1 and OsNramp5 Mutants Generated Through CRISPR/Cas9-Mediated Mutagenesis for Breeding Low Cd Rice. Rice Sci. 2019, 26, 88–97. [Google Scholar] [CrossRef]
  259. Chang, J.D.; Huang, S.; Yamaji, N.; Zhang, W.; Ma, J.F.; Zhao, F.J. OsNRAMP1 Transporter Contributes to Cadmium and Manganese Uptake in Rice. Plant Cell Environ. 2020, 43, 2476–2491. [Google Scholar] [CrossRef]
  260. Chu, C.; Huang, R.; Liu, L.; Tang, G.; Xiao, J.; Yoo, H.; Yuan, M. The Rice Heavy-Metal Transporter OsNRAMP1 Regulates Disease Resistance by Modulating ROS Homoeostasis. Plant Cell Environ. 2022, 45, 1109–1126. [Google Scholar] [CrossRef] [PubMed]
  261. 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]
  262. Nieves-Cordones, M.; Mohamed, S.; Tanoi, K.; Kobayashi, N.I.; Takagi, K.; Vernet, A.; Guiderdoni, E.; Périn, C.; Sentenac, H.; Véry, A.A. Production of Low-Cs+ Rice Plants by Inactivation of the K+ Transporter Os HAK 1 with the CRISPR-Cas System. Plant J. 2017, 92, 43–56. [Google Scholar] [CrossRef] [Green Version]
  263. Derksen, D.A.; Anderson, R.L.; Blackshaw, R.E.; Maxwell, B. Weed Dynamics and Management Strategies for Cropping Systems in the Northern Great Plains. Agron. J. 2002, 94, 174–185. [Google Scholar] [CrossRef]
  264. Riaz, M.; Jamil, M.; Mahmood, T.Z. Yield and Yield Components of Maize as Affected by Various Weed Control Methods Under Rain-Fed Conditions of Pakistan. Int. J. Agric. Biol. 2007, 9, 152–155. [Google Scholar]
  265. Chipman, D.; Barak, Z.; Schloss, J.V. Biosynthesis of 2-aceto-2-hydroxy Acids: Acetolactate Synthases and Acetohydroxyacid Synthases. Biochim. Biophys. Acta 1998, 1385, 401–419. [Google Scholar] [CrossRef] [PubMed]
  266. Lee, K.Y.; Townsend, J.; Tepperman, J.; Black, M.; Chui, C.F.; Mazur, B.; Dunsmuir, P.; Bedbrook, J. The Molecular Basis of Sulfonylurea Herbicide Resistance in Tobacco. EMBO J. 1988, 7, 1241–1248. [Google Scholar] [CrossRef] [PubMed]
  267. Sun, Y.; Zhang, X.; Wu, C.; He, Y.; Ma, Y.; Hou, H.; Guo, X.; Du, W.; Zhao, Y.; Xia, L. Engineering Herbicide-Resistant Rice Plants Through CRISPR/Cas9-Mediated Homologous Recombination of Acetolactate Synthase. Mol. Plant 2016, 9, 628–631. [Google Scholar] [CrossRef] [Green Version]
  268. Tian, S.; Jiang, L.; Cui, X.; Zhang, J.; Guo, S.; Li, M.; Zhang, H.; Ren, Y.; Gong, G.; Zong, M.; et al. Engineering Herbicide-Resistant Watermelon Variety Through CRISPR/Cas9-Mediated Base-Editing. Plant Cell Rep. 2018, 37, 1353–1356. [Google Scholar] [CrossRef]
  269. Svitashev, S.; Young, J.K.; Schwartz, C.; Gao, H.; Falco, S.C.; Cigan, A.M. Targeted Mutagenesis, Precise Gene Editing, and Site-Specific Gene Insertion in Maize Using Cas9 and Guide RNA. Plant Physiol. 2015, 169, 931–945. [Google Scholar] [CrossRef] [Green Version]
  270. Kuang, Y.; Li, S.; Ren, B.; Yan, F.; Spetz, C.; Li, X.; Zhou, X.; Zhou, H. Base-Editing-Mediated Artificial Evolution of OsALS1 In Planta to Develop Novel Herbicide-Tolerant Rice Germplasms. Mol. Plant 2020, 13, 565–572. [Google Scholar] [CrossRef]
  271. Wang, F.; Xu, Y.; Li, W.; Chen, Z.; Wang, J.; Fan, F.; Tao, Y.; Jiang, Y.; Zhu, Q.-H.; Yang, J. Creating a Novel Herbicide-Tolerance OsALS Allele Using CRISPR/Cas9-Mediated Gene Editing. Crop J. 2021, 9, 305–312. [Google Scholar] [CrossRef]
  272. Schönbrunn, E.; Eschenburg, S.; Shuttleworth, W.A.; Schloss, J.V.; Amrhein, N.; Evans, J.N.; Kabsch, W. Interaction of the Herbicide Glyphosate with Its Target Enzyme 5-Enolpyruvylshikimate 3-Phosphate Synthase in Atomic Detail. Proc. Natl. Acad. Sci. USA 2001, 98, 1376–1380. [Google Scholar] [CrossRef] [Green Version]
  273. Ortega, J.L.; Rajapakse, W.; Bagga, S.; Apodaca, K.; Lucero, Y.; Sengupta-Gopalan, C. An Intragenic Approach to Confer Glyphosate Resistance in Chile (Capsicum annuum) by Introducing an In Vitro Mutagenized Chile EPSPS Gene Encoding for a Glyphosate Resistant EPSPS Protein. PLoS ONE 2018, 13, e0194666. [Google Scholar] [CrossRef] [Green Version]
  274. 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]
  275. Zhang, R.; Liu, J.; Chai, Z.; Chen, S.; Bai, Y.; Zong, Y.; Chen, K.; Li, J.; Jiang, L.; Gao, C. Generation of Herbicide Tolerance Traits and a New Selectable Marker in Wheat Using Base Editing. Nat. Plants 2019, 5, 480–485. [Google Scholar] [CrossRef] [PubMed]
  276. Zhang, R.; Chen, S.; Meng, X.; Chai, Z.; Wang, D.; Yuan, Y.; Chen, K.; Jiang, L.; Li, J.; Gao, C. Generating Broad-Spectrum Tolerance to ALS-Inhibiting Herbicides in Rice by Base Editing. Sci. China Life Sci. 2021, 64, 1624–1633. [Google Scholar] [CrossRef] [PubMed]
  277. Li, Z.; Liu, Z.B.; Xing, A.; Moon, B.P.; Koellhoffer, J.P.; Huang, L.; Ward, R.T.; Clifton, E.; Falco, S.C.; Cigan, A.M. Cas9-Guide RNA Directed Genome Editing in Soybean. Plant Physiol. 2015, 169, 960–970. [Google Scholar] [CrossRef] [Green Version]
  278. 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]
  279. Liu, L.; Kuang, Y.; Yan, F.; Li, S.; Ren, B.; Gosavi, G.; Spetz, C.; Li, X.; Wang, X.; Zhou, X.; et al. Developing a Novel Artificial Rice Germplasm for Dinitroaniline Herbicide Resistance by Base Editing of OsTubA2. Plant Biotechnol. J. 2021, 19, 5–7. [Google Scholar] [CrossRef] [PubMed]
  280. Mao, X.; Zheng, Y.; Xiao, K.; Wei, Y.; Zhu, Y.; Cai, Q.; Chen, L.; Xie, H.; Zhang, J. OsPRX2 Contributes to Stomatal Closure and Improves Potassium Deficiency Tolerance in Rice. Biochem. Biophys. Res. Commun. 2018, 495, 461–467. [Google Scholar] [CrossRef]
Figure 1. Applications of genome editing in crop improvement against abiotic stresses.
Figure 1. Applications of genome editing in crop improvement against abiotic stresses.
Life 13 01456 g001
Figure 2. Mechanism of genome editing using CRISPR/Cas9.
Figure 2. Mechanism of genome editing using CRISPR/Cas9.
Life 13 01456 g002
Table 5. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for developing salinity tolerance.
Table 5. List of reported targeted gene(s) via CRISPR/Cas9 technology in different plant species for developing salinity tolerance.
CropsGeneTraitTechniqueReferences
RiceOsbHLH024SalinityCRISPR/Cas9[235]
RiceOsRR22SalinityCRISPR/Cas9[232,237]
RiceOsRAV2, OsNAC041, OsmiR535SalinityCRISPR/Cas9[234,236,238]
RiceOsRR9, OsRR10SalinityCRISPR/Cas9[239]
RiceOsNAC041SalinityCRISPR/Cas9[240]
RiceOsOTS1SalinityCRISPR/Cas9[241,242]
RiceOsDSTDrought and salinityCRISPR/Cas9[151]
RiceSAPK2Tolerance to salinityCRISPR/Cas9[131]
WheatTaHAG1Salt toleranceCRISPR/Cas9[243]
MaizeZmHKTITolerance to salinityCRISPR/Cas9[244]
SoybeanGmAITRSalt toleranceCRISPR/Cas9[245]
SoybeanDrb2a, Drb2bTolerance to droughtand salinity stressCRISPR/Cas9[208]
BarleyHvITPK1salinityCRISPR/Cas9[246]
TomatoSlHyPRP1, SlARF4salinityCRISPR/Cas9[247,248]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yadav, R.K.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Asati, R.; Chauhan, S.; Tiwari, P.N.; Payasi, D.K. Genome Editing and Improvement of Abiotic Stress Tolerance in Crop Plants. Life 2023, 13, 1456. https://doi.org/10.3390/life13071456

AMA Style

Yadav RK, Tripathi MK, Tiwari S, Tripathi N, Asati R, Chauhan S, Tiwari PN, Payasi DK. Genome Editing and Improvement of Abiotic Stress Tolerance in Crop Plants. Life. 2023; 13(7):1456. https://doi.org/10.3390/life13071456

Chicago/Turabian Style

Yadav, Rakesh Kumar, Manoj Kumar Tripathi, Sushma Tiwari, Niraj Tripathi, Ruchi Asati, Shailja Chauhan, Prakash Narayan Tiwari, and Devendra K. Payasi. 2023. "Genome Editing and Improvement of Abiotic Stress Tolerance in Crop Plants" Life 13, no. 7: 1456. https://doi.org/10.3390/life13071456

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