Next Article in Journal
MiR-93/miR-375: Diagnostic Potential, Aggressiveness Correlation and Common Target Genes in Prostate Cancer
Next Article in Special Issue
CRISPR/Cas9-Mediated Generation of Pathogen-Resistant Tomato against Tomato Yellow Leaf Curl Virus and Powdery Mildew
Previous Article in Journal
A Practical Review of NMR Lineshapes for Spin-1/2 and Quadrupolar Nuclei in Disordered Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Revolution toward Gene-Editing Technology and Its Application to Crop Improvement

1
College of Plant Sciences and Technology Huazhong Agricultural University, Wuhan 430070, China
2
Institute of Biological Sciences, University of Talca, 2 Norte 685, Talca 3460000, Chile
3
Key Laboratory of Arable Land Conservation (Middle and Lower Reaches of Yangtze River), Ministry of Agriculture, College of Resources and Environment, Huazhong Agricultural University, Wuhan 430070, China
4
Sate Key Laboratory of Agricultural Microbiology and State Key Laboratory of Microbial Biosensor, College of Life Sciences Huazhong Agriculture University Wuhan, Wuhan 430070, China
5
Department of Plant Pathology, University College of Agriculture and Environmental Sciences, The Islamia University of Bahawalpur, Bahawalpur 63100, Pakistan
6
National Institute for Biotechnology and Genetic Engineering (NIBGE), Faisalabad 38000, Pakistan
7
Department of Bioscience, COMSATS Institute of Information Technology, Islamabad 45550, Pakistan
8
Graduate School of Biotechnology & Crop Biotech Institute, Kyung Hee University, Yongin 17104, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(16), 5665; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21165665
Submission received: 14 July 2020 / Revised: 4 August 2020 / Accepted: 5 August 2020 / Published: 7 August 2020

Abstract

:
Genome editing is a relevant, versatile, and preferred tool for crop improvement, as well as for functional genomics. In this review, we summarize the advances in gene-editing techniques, such as zinc-finger nucleases (ZFNs), transcription activator-like (TAL) effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR) associated with the Cas9 and Cpf1 proteins. These tools support great opportunities for the future development of plant science and rapid remodeling of crops. Furthermore, we discuss the brief history of each tool and provide their comparison and different applications. Among the various genome-editing tools, CRISPR has become the most popular; hence, it is discussed in the greatest detail. CRISPR has helped clarify the genomic structure and its role in plants: For example, the transcriptional control of Cas9 and Cpf1, genetic locus monitoring, the mechanism and control of promoter activity, and the alteration and detection of epigenetic behavior between single-nucleotide polymorphisms (SNPs) investigated based on genetic traits and related genome-wide studies. The present review describes how CRISPR/Cas9 systems can play a valuable role in the characterization of the genomic rearrangement and plant gene functions, as well as the improvement of the important traits of field crops with the greatest precision. In addition, the speed editing strategy of gene-family members was introduced to accelerate the applications of gene-editing systems to crop improvement. For this, the CRISPR technology has a valuable advantage that particularly holds the scientist’s mind, as it allows genome editing in multiple biological systems.

1. Introduction

The rapidly growing population and a wide range of competitive dairy products and meat are pushing agricultural output and expanding the demand for feed, food, biofuels, and livestock [1]. By 2050, the worldwide population will expand up to >9 billion, which may boost crop production demand by 100–110%. Consequently, the effective production of staple crops, such as Oryza sativa (rice), Triticum aestivum (wheat), Zea mays (maize), and Glycine max (soybean), will increase by just 38–67% [1,2]. Currently, numerous genome-editing tools and techniques have been adopted to overcome the problems arising in plants to compensate for the increased demand for food in the future [3]. Gene-editing techniques, such as engineered endonucleases/meganucleases (EMNs), zinc-finger nucleases (ZFNs), TAL effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR) [4], are important tools in plant research, as they allow the remodeling of future crops.
ZFNs were the first truly targeting protein reagents to revolutionize the genome manipulation area of research. ZFNs are binding domains for DNA that recognize three base pairs at the target site [5]. ZFNs have been commonly used for targeted genome modification in different plant species, such as Arabidopsis thaliana (Arabidopsis), Nicotiana tabacum (tobacco), and maize [6,7,8]. Another site-driven mutagenesis genome-editing system, TALENs, was defined first in plant pathogenic bacteria (Xanthomonas) and is based on a concept similar to that of ZFNs. TALENs target one nucleotide at the target site (instead of three), thus rendering TALENs precise [9]. TALENs were successfully used for genome editing in angiosperms and bryophytes [10,11].
Extensive investigation in this field led to the development of new genome-editing tools, such as CRISPR/Cas9 and CRISPR/Cpf1 [12,13]. Initially, these techniques were developed in prokaryotes, because there were no efficient genome-editing techniques for eukaryotes at specific sites. However, at the advent of eukaryotic genome editing, the CRISPR technology has revolutionized our ability to generate specific changes in crops [14]. The CRISPR system requires only the guide RNA sequence to be changed for each DNA target site. Under different circumstances, the usage and modification of CRISPR technology are quite simple and efficient [15,16]. In this review, we highlight the use of genome-editing techniques to achieve highly precise and desired modifications in plants, as well as examples of the application of EMNs, ZFNs, TALENs, and CRISPR/Cas9/Cpf1 in various plants (Figure 1).

2. The Journey from Engineered Meganucleases (EMNs) to CRISPR

2.1. Meganucleases (MNs)

Among endonucleases, meganucleases are characterized by the presence of a broad recognition site of about 12–40 bp. Because of their specific nature and long recognition site, these enzymes are regarded as the most precise restriction enzymes [17]. Therefore, meganucleases are also known as homing endonucleases. Repair in double-stranded breaks (DSBs) occurs via nonhomologous end joining (NHEJ), which is functionally responsible for knocking out genes in tobacco and Arabidopsis plants [18,19]. However, it is difficult to remodify meganucleases together with other genome-targeting techniques, because DNA-binding domains are often intermingled with the catalytic domain of meganucleases and cannot be detached from one another [20]. The outcomes of previous research demonstrated gene editing in plants by incorporating modified meganucleases in Arabidopsis, Gossypium hirsutum (cotton), and maize. However, additional efforts are needed to improve this approach, as the manipulation of meganucleases seems to be difficult. Hence, researchers focused on other more efficient, accurate, and simpler methods of gene editing, such as ZFNs, TALENs, and CRISPR.

2.2. Zinc-Finger Nucleases (ZFNs)

Zinc-finger nucleases (ZFNs) are one of the most efficient and effective tools for genome editing by targeting DSBs [21]. The first generation of genome-editing techniques based on ZFNs was developed using chimerically engineered nucleases. This approach was enabled by the discovery of the functional Cys2-His2 zinc-finger domain [4,22]. Fundamentally, the structural composition of ZFNs involves two domains: (1) The DNA-binding domain, which consists of 300–600 zinc-finger repeats [23]. Each zinc-finger repeat can monitor and read between 9 and 18 base pairs (bp); and (2) the DNA cleavage domain, which is known as the nonspecific cleavage domain of the type II restriction endonuclease Fok1 and acts as the DNA cleavage domain in ZFNs [24]. ZFNs contain two monomers attributed to their respective target sequences reversely flanking in between 5 and 6 bp of the DNA target [24,25]. Dimers containing Fok1 domains slice DNA within its flanking sequence (Figure 2). The specific sequence of 24–30 bp is monitored by a zinc-finger domain that has specific or rare targeting sites in the genome [26]. The field of genome editing is progressing by acquiring the ability to engineer and manipulate applied and basic genomic targets.

2.3. Transcriptional Activator-Like Effector Nucleases (TALENs)

The TALEN system for accurate genome editing is a commonly adopted method that has been in use for several years [27]. TALENs were developed via the amalgamation of the FokI cleavage domain with the DNA-binding domains of TALE proteins. TALEs comprise multiplex repeats of 34 amino acids for the efficient edition of a single base pair [28]. Like ZFNs, TALENs also promote targeted DSBs that help initiate pathways that are responsible for DNA damage and ensure modifications [4]. The proteins involved in the TALEN system comprise a central domain, which is responsible for binding to DNA, and a nuclear localization sequence [29] (Figure 3). In 2007, it was observed for the first time that these proteins possess the capability of binding to DNA. However, the DNA-binding domain includes a 34-amino-acid repeated sequence, with each repeat perceiving a single nucleotide in the target DNA, whereas each repeated sequence of ZFNs perceived three nucleotides in the target DNA [30].
The number of studies that used ZFNs and TALENs in plants is comparatively small, and these reports appear to favor TALENs; however, the efficiency of editing achieved by these two nucleases is quite low. Therefore, the use of TALENs is more unaffected and favorable for programming. The targets of TALEs are specifically recognized by the occurrence of repeat variable di-residue (RVD) flanking at 12 and 13 positions of each target sequence [26,27].
Generally, TALE proteins can be modulated by binding DNA repeated sequences. Previous studies showed that the nucleotides of the DNA sequence are fixed by the help of TALE proteins always at the 5′ end thymidine base. In the absence of a 5′T, the activities of TALE transcription factors (TALE-TFs) and TALE recombinase (TALE-R) are decreased [31]. TALENs are preferred over ZFNs because their modulation is much simpler and cost effective, with a much lower off-target rate (Table 1).

2.4. CRISPR/Cas9

This genome-editing technique, which relies on the activity of RNA-guided nucleases and their mode of action, has gained much attention because of its versatility, potency, adequacy, and simplicity [32]. The CRISPR/Cas9 system is a highly conserved system that originated from the bacterial species Streptococcus pyogenes [33,34]. Its discovery was a significant breakthrough of the 20th century, as it represented an entirely distinct and divergent tool that was quickly examined by many bioinformaticians, biotechnologists, and microbiologists.
In the 2012–2013 period, the CRISPR/Cas9 system was successfully implemented with remarkable cutting efficiency and simplicity to modify animal and plant genes [35,36]. Studies reported three CRISPR/Cas systems (I, II, and III), each of which has distinct molecular mechanisms for nucleic acid piercing and targeting [33,37]. The initial identification of Cas9 (formerly known as COG3513, Csx12, Cas5, or Csn1) through bioinformatics analyses revealed that it acts as a large multifunctional protein structure that comprises two nuclease domains, HNH and RuvC-like [38]. The development of the CRISPR system proved to be advantageous for the manipulation of genetically modified cells in living organisms, as well as in culture (Figure 4) [39].
Because of its versatility, simplicity, efficacy, and wide range of applications, the CRISPR/Cas9 system has been applied in many fields of research, such as biotechnology, genetic engineering, and fundamental and applied biology (Table 1).
With the expansion of the plant genome-editing system, the expression cassette of CRISPR/Cas9 is transformed into the cells, incorporated into the nuclear genome, and expressed, followed by the cleavage of its target DNA sequence, usually 3 bp upstream of the protospacer adjacent motif (PAM) site. Double-stranded breakage of DNA activates two separate mechanisms of DNA repair, NHEJ and homology-directed repair (HDR) [40]. In the absence of a homologous template, NHEJ mediates the direct re-ligation of the broken DNA molecules, normally leading to insertions and deletions (InDels), or substitutions at the DSB site. However, in the presence of a donor DNA sequence, HDR may add new alleles, correct existing changes, or insert new sequences of interest [15,41]. Although DNA becomes integrated into the plant genomic site at a low frequency [42], the integrated transgene can still be expressed and becomes functional only for a short period. Therefore, the expression of CRISPR/Cas9 via transgenesis may offer an alternative method for genome editing in plants. Interestingly, two simple and effective methods adopted for genome editing rely on the expression profile of the CRISPR/Cas9 DNA or RNA [43]. For these methodologies, antibiotic and herbicide selection steps are adopted during post-transformation tissue culture and obstructed, which yield plants that regenerate from the induction cells of the callus that functionally express the CRISPR/Cas9 system.

2.5. New Tools for Plant Genome Editing

Based on the revolution of molecular biology and the discovery of sequences in the microbial immune system, biotechnologists are now able to manipulate the genome of organisms of interest in a specific and precise way with the aid of CRISPR and its associated Cas proteins. This remarkable genome-editing system is categorized into two broad classes and six subtypes. CRISPR class II has a type V effector termed Cpf1, which can be designed using highly specific CRISPR RNA to cleave the corresponding DNA sequences [44,45]. Cpf1 has various distinct features, such as the ability to target T-rich motifs, the absence of a requirement for trans-activating crRNA, the versatile capacity to induce a staggered double-strand break, and the potential for both RNA processing and DNA nuclease activity; hence, it represents an alternative to Cas9 [46].
The Cpf1 nuclease or Cas12a was recently discovered in Prevotella and Francisella1 at the MIT and the Broad Institute (USA) by Zhang and his team [47]. Regarding its structural configuration, cpf1 belongs to type V among the CRISPR systems and is a monomeric protein comprising 1200–1500 amino acids. It recognizes a 5′-TTTN-3′ or 5′-TTTV-3′ sequence (V = A, C, or G), in some cases as PAM in a DNA sequence, and the whole array consists of nine spacer sequences, which are disassociated with 36-nucleotide-long repeated sequences (Table 1) [48], ultimately leading to a spacer derived from a part of the crRNA that is complementary to the target DNA [49]. One of the unique features that render Cpf1 a highly useful nuclease is the formation of staggered ends. It contains five 5–8-nucleotide-long overhangs depending on the crRNA length at the site of cleavage [50]. These overhangs enable genome manipulation and provide a flexible approach for base editing and epigenetic modulation [51].
Recent reports suggest that Cpf1 can cleave double-stranded DNA at a single catalytic site in the RuvC domain, whereas the Nuc domain is responsible for the regulation of the substrate DNA [52] (Figure 5). Another report suggests that small molecular compounds can enhance the efficiency of Cpf1, as they are directly involved in activating or suppressing signaling pathways for cellular repair. Thus, small-molecule-mediated DNA repair aids CRISPR-mediated knockout strategies [53].
Furthermore, many desirable traits can only be obtained in crops by correctly inserting or removing segments of DNA. Base editing provides a new method for base substitution. However, the conversions of C–T and A–G remain limited [53,54]. Recently, a groundbreaking genome editor, “prime editing,” was developed that can directly insert new genetic information into a designated DNA site, thus dramatically expanding the genome edition range and capabilities [55]. Cas9 is a nickase fused with reverse transcriptase in the prime editing system, and sgRNA is replaced by the prime editing guide RNA (pegRNA), which includes the target site identification of sgRNA and the RNA template specifying the DNA sequence for insertion on the target genome [55].

2.6. Applications

2.6.1. MNs, ZFNs, and TALENs

The use of ZFNs was primarily examined in Arabidopsis [19,72] and tobacco as model plants. Because of the highly specific nature of the engineered nucleases used for the targeting of genes, researchers believed that this technique with further modifications could be applied to other crop plants [73]. After a few years, many studies reported the results of ZFN-mediated gene-targeting approaches in many crops, including tobacco [18], maize [74], and model plants (Arabidopsis) [75]. The application of ZFNs for gene targeting in various plant species is presented in the following sections (Table 2). The only drawback of ZFNs is their ability to bind to any nucleotide sequence (one zinc finger can bind to three nucleotides in the targeted DNA), as well as their ability of binding to off-target sequences [23,76]. Genome editing via ZFNs was also achieved in soybean by targeting DICER-like (DCL) genes. The results showed that mutation is comprehensively effective for the transmission of inheritance due to ZFN-induced mutation. The context-dependent assembly scaffold is an open and rapid method that is used for modulating novel ZFN arrays [77].
In addition, findings from previous studies suggest a method for targeted mutagenesis in the genome of paleopolyploid soybean using ZFN that is competent in targeting single or multiple copies of gene families [78]. It has also been reported that the ZFN protein activated the transcriptional machinery of the b-ketoacyl-ACP Synthase II gene in Brassica napus [79]. Interestingly, engineered ZFN-TFs can play a significant role in the modification of agronomic traits, as well as endogenous genes [80]. Previous outcomes demonstrated the importance of artificially engineered TALEs, thus allowing the use of TALE-binding code for DNA targeting sites together with TALE DNA-binding domains (DBDs).
Therefore, DBD can be amalgamated with an effector or catalytic domain-like nuclease, e.g., a nuclease, to achieve a remarkable tool for DNA editing [81,82,83]. TALE-fusion proteins employ the C-terminal region of the central repeat domain, which acts as a linker between TALE DBD and the effector domain. The linker length may vary according to the effector domain that is used for the dimerization of the FoKI nuclease domain [84,85,86]. However, the length of the longer linker used for the activation of the domain is 65 amino acids. The presence of the DELLA gene in tomato, known as PROCERA (PRO), yielded an inhibitory effect on the regulation of the signaling cascade of gibberellic acid, whereas TALEN edited the PRO gene under the control of an estrogen-inducible promoter, resulting in phenotypes with a consistently increased GA response [87].

2.6.2. CRISPR/Cas9 and CRISPR/Cpf1

CRISPR/Cas9 and CRISPR/Cpf1 bring a revolutionary change into the field of biology, and many laboratories around the world are adopting this leading-edge technology because of its tremendous applications. In this section, we summarize the advantages of this powerful approach to engineer genes and their functions for crop improvement.

Improvement in Yield and Quality via CRISPR/Cas9 and CRISPR/Cpf1

Genome-editing technologies have far-reaching large-scale practical applications to overcome one of the key milestones of modern biotechnology, i.e., the development of new crop varieties with high yield, resistance to biotic and abiotic stresses, and high nutritional value. Much progress has been achieved using the CRISPR technology. Recently, an oil known as “biotech oil” was obtained from Camelina sativa seeds that has wide applications with an enhanced fatty acid composition. It is not only beneficial for human health because of its potency to resist oxidation but also applicable for the production of chemicals that are synthesized commercially, such as biofuels [97,98]. Genomic editing using CRISPR/Cas9-mediated technology has also been used in woody species, such as poplar (Populus tomentosa Carr). The phytoene desaturase 8 (PtoPDS-8) gene was edited in a site-specific manner using four sgRNAs. A phenotypic analysis of transgenic poplar plants showed an albino phenotype with about 51% of induced mutation frequency [99]. The 4-coumarate: CoA ligase-1 and 4-coumarate: CoA ligase-2 (4CL1 and 4CL2) genes also participate in the synthesis of lignin and flavonoid in poplar plants. Using CRISPR/Cas9, the 4CL1 and 4CL2 gene families were mutated under the control of the ubiquitin-6 (U6) promoter of Medicago, resulting in the generation of a bi-allelic mutation with a 100% efficiency. The homologs and multiplex recombination-mediated editing of the arabidopsis phytoene desaturase (PDS3) gene were obtained with a measured frequency of 65–100% [100,101].
Interestingly, Zhang et al. (2016) [102] engineered a multiple CRISPR/Cas9 system that has been used for fast editing and observed the presence of six editing PYL gene families of ABA receptors with a 13–93% mutation frequency in the T1 generation in a single transformation experiment. The brassinosteroid insensitive 1 (BRI1), jasmonate-zim-domain protein 1 (JAZ1), and gibberellic acid insensitive genes were also engineered using CRISPR-Cas9, with a mutation frequency of 26–84% [42,69]. The flowering Locus T (FT) and squamosa promoter-binding protein-like 4 genes were also edited with CRISPR/Cas9; 90% of plants in the T1 generation carried a mutation in the late flowering stage [100]. Mutagenesis of the green fluorescent protein gene in Nicotiana benthamiana using Cas9 RNA-guided endonuclease has been ameliorated [103]. This system was transformed using a tobacco rattle virus vector to modulate the regulation of plant genes via engineering and to edit transcriptional factors [104].
The generation of homozygous rice was edited using the CRISPR/Cas9 system by Zhang (2014) and Zhou (2014) [105,106]. The results of these studies suggest that deletions in the chromosomal gene cluster, as well as small heritable variations in the genetic makeup during the genome editing by CRISPR-Cas9, were present in T0 plants. CRISPR/Cas9-mediated mutagenesis was also examined in three members of the rice aldehyde oxidase (AOX1) gene family (OsAox1a, OsAox1b, and OsAox1c) and in the OsBEL protein of rice; moreover, the inherent modification of the transgene in the next generation was also reported by Xu et al. [107]. The barley HvPM19 gene encodes an ABA-inducible membrane protein that is involved in the upregulation of grain dormancy. Mutation induced by Cas9 in two copies of HvPM19 yielded a 10% mutation frequency [45,108].
It has been acknowledged that the most extensively used wild-type spCas9 is vigorous in identifying both NGG and NAG PAMs in rice. Other applications for producing high-quality crops using an efficient CRISPR-Cas9 system include seeds with a high concentration of oleic acid oil in Camelina sativa and B. napus, and the targeting of ALCATRAZ genes to enhance pod shattering in B. napus [109,110]. CRISPR technology is optimal to generate targeted gene knockouts. However, several essential genes cause seedling lethality when knocked out, and several agronomic traits (such as improved photosynthesis) require gene overexpression [111]. The roles of the grain number (Gn1a) and grain size (GS3) QTLs were investigated with the help of a CRISPR/Cas9-mediated QTL-editing approach in rice [112].
The CRISPR/Cas9 technique was used mainly as a proof of concept in many vegetable crops, e.g., in cabbage, Chinese kale, and watermelon, to induce mutations in the PDS gene [113,114]. In the case of vegetables, CRISPR/Cas9 studies have been performed most frequently in tomatoes, because of the economic value of the crop or the ease of genetic transformation using agrobacterium. Parthenocarpy can be a desirable trait in tomatoes because of consumer preference and treatment purposes [115]. Soyk et al. (2017) [116] indicated that the targeted mutagenesis of the engineered self-pruning 5G (SP5G) gene of tomato yielded early flowering and more bush, which in turn resulted in an early harvest. Brooks et al. (2014) [117] magnified the CRISPR/Cas9-mediated obstruction of the ARGONAUTE 7 (SlAGO7) gene and observed a needle-like or wiry leaf phenotype in tomato (Table 3). The MADS-box transcription factor-encoding RIPENING INHIBITOR (RIN) gene was found to regulate fruit ripening in tomato. This technique was modulated to engineer three target regions within the gene; RIN mutant (homozygous) tomato plants displayed incomplete ripening with a low pigmentation (red) rate compared with wild-type plants, which demonstrated the crucial role of RIN in the ripening process [100,118]. Furthermore, orthologs of GA4 in B. oleracea, BolC.GA4.a, were used to induce 10% targeted mutations by Cas9 and led to a dwarf phenotype that was linked with GA4 knockout [71,100,119,120].
Furthermore, the biosynthesis of steroidal glycoalkaloids (SGAs) in potato was used together with the CRISPR/Cas9 method to target 16α-hydroxylase steroids (St16DOx). This research provided two SGA-free potato lines with deletions of St16DOX [121]. Similarly, the starch synthase GBSS gene has been mutated in potato via CRISPR/Cas9. The mutated lines showed reduced amylose levels and an increased concentration of the amylose/amylopectin ratio [122]. Furthermore, the SnLazy1 locus, which is the tomato ortholog of Lazy1, was edited by CRISPR/Cas9 in Solanum nigra, with successful inheritance of the removal of two separate snlazy1-cr alleles and the production of plants with stem development in a relatively downward direction [123].
Genome editing may be the only way to improve this important staple food and fruit. To date, only a small number of fruit-producing species (citrus, tomatoes, watermelons, grapes, or strawberries) with traits inherited from CRISPR/Cas9 via the germline have been recorded [118]. Genome editing of gibberellin biosynthesis has allowed the generation of dwarf fruit trees [124], with the capacity for a high productivity rate through dense planting and decreased usage of water and fertilizers and lower land and labor costs. Moreover, genome editing for the inhibition of ethylene biosynthesis was found to play an essential role in the fruit-ripening process [125]. Moreover, its signaling pathways led to the development of new varieties with an increased shelf life [93]. The findings of previous research consistently presented a novel Xanthomonas citri, which expedited the technique of agroinfusion to transfer CRISPR-Cas9 for targeting the CsPDS gene into sweet orange leaves [119].
Because of the high efficacy of genome editing, which does not allow the involvement of foreign DNA, it will be easy for the consumer to utilize genome-edited fruits. A useful prediction was made by a researcher [120,121], who recently determined that gene editing in golden apple protoplasts can be accomplished by adopting similar Cas9/sgRNA RNP complexes, as discussed earlier for genome editing in other crops. Another group investigated a wild-type species of tomato called groundcherry (Physalis pruinosa) that produces a high yield of large fruits [39]. In addition, the PPO mutation in apples can be considered transgene free using CRISPR/Cas9 and could easily be applicable worldwide [122]. CRISPR/Cas9 is committed to the development of seedless fruits through the modification/mutation of genes responsible for seed formation. In tomatoes, parthenocarpy has also been recorded by knockout or mutation of the SIAGL6 and SIIAA9 genes by CRISPR/Cas9 (Table 3) [123]. The parthenocarpy production method controlled by CRISPR/Cas9 can be implemented in fruits, such as citrus, custard apple, grapes, kinnow, peach, and watermelon, among which there is a high demand for seedless fruit. CRISPR/Cas9 has also recently been used to cause mutations in the MaGA20ox2 gene, which regulates banana dwarfism [126].
The CRISPR/Cpf1 method has been used to edit the FAD2-1B and FAD2-1A genes to enhance the oil composition of soybean to produce high-yielding soybean plants with higher oleic acid levels [127]. Using CRISPR/Cpf1, plant breeders have correctly improved production and quality with a high degree of effectiveness [128]. Various Cpf1 proteins were used to mediate the editing of the genomes among various higher plant species, such as tobacco, soybean, and rice. In recent years, the OsPDS and OsBEL genes were targeted by Cpf1 and were engineered by selecting two genomes within rice for stability and heritage mutations [124,125]. The Chlorophyllidea oxygenase (CAO1) gene, which converts chlorophyll a into chlorophyll b, has been targeted for gene insertion in rice using CRISPR/Cpf1 [67,124]. Gene editing using the guide gRNA-Cas9/Cpf1 ribonucleoprotein (RNP) is suitable for fruit tree protoplasts that have been displayed in apple and grape cells [120].

Upgrading of Climate-Resilient Crops, Vegetables, and Fruits

The CRISPR technology is widely used together with a variety of biotic and abiotic stresses in major crop plants, such as wheat, rice, corn, cotton, soybeans, tomato, and potato. The CRISPR tool has modernized plant breeding programs for the production of smart climate abiotic stress-tolerant crops. Moreover, CRISPR/Cas9 is a novel technique that can be used to knock out the eukaryotic translational initiation factor eIF4E gene, which is necessary for the translation process in vegetables, such as Cucumis sativus. The resultant gene knockout ensures resistance against viruses, such as the papaya ringspot mosaic virus-W (PRSV-W), the zucchini yellow mosaic virus (ZYMV), and the cucumber vein yellowing virus (CVYV) [148].
Rice production is significantly decreased by high levels of salt in the soil. The mechanism of salt tolerance in rice was determined by CRISPR/Cas9. A CRISPR/Cas9-mediated knockout mutant of the OsPRX2 gene exhibited a higher level of antioxidant induction compared with the usual reactive oxygen species (ROS) accumulation [149]. ROS perform an important role in plants by acting as signaling molecules for gene expression regulation, viral pathogen protection, and the symbiotic fixation of nitrogen between the plant and soil rhizobia [150,151].
The CRISPR-Cas9 advanced breeding technology has enabled the development of a new ARGOS8 variant in maize. Compared with wild-type alleles, the ARGOS8 variant showed an improved grain yield under flowering stress conditions (five bushels per acre). These findings demonstrated that the CRISPR/Cas9 system is an accurate tool for the generation of new allelic variations in crops for the growth of drought-resistant plants [152]. Furthermore, the knockout of two genes Drb2a and Drb2b via CRISPR/Cas9 identified their role in controlling salt and drought tolerance in soybean [153]. In tomatoes, molecules, such as mitogen-activated protein kinases, which are responsible for drought stress under the protection of membrane cells against oxidation and via the regulation of transcription genes to manage dry stress, are significant signals. The control of the drought tolerance mechanism via the SIMAPK3 gene was reported in the tomato system, which produces knockout mutants of the SlMAPK3 gene under dry stress using CRISPR/Cas9 (Table 4) [154].
Recent studies reported the editing of vegetable crops using CRISPR/Cas9 for the development of different valuable traits. The CRISPR-Cas9-mediated genome editing for heat tolerance was achieved by targeting the SlAGAMOUS-LIKE 6 (SIAGL6) gene in tomato. Knockout of the SIAGL6 gene improved the fruit setting of tomato under heat stress [123]. Similarly, another group used a different approach to develop knockout mutations in the SIIAA9 gene, which were involved in the auxin signaling pathway to repress the initiation of the development of fruits without fertilization. In addition, in other horticultural plants, such as watermelon, bitter gourd, amber gourd, etc., seedless fruits or fruits with less seeds can be obtained using this precise and rapid method of developing parthenocarpy [115]. In grapes, knockout of WRKY52, which encodes a transcription factor related to biotic stress responses, via CRISPR/Cas9 enhanced resistance to Botrytis cinerea [155].
In the presence of 100 mM NaCl, self-pollinated offspring tomato plants, which bear the HKT1;2 HDR allele, exhibited stable inheritance germination tolerance. Transgene-free edited plants that reproduce asexually and sexually have been developed using CRISPR/Cpf1 [156]. Hence, the studies mentioned above revealed that CRISPR/Cas9 plays an important role in the development of climate-resilient crops, vegetables, and fruits.

Application of CRISPR/Cas9 and CRISPR/Cpf1 to Plant Disease Resistance

Recent advances have been made that cover the major area of genome-editing applications in plant breeding to generate varieties that are resistant against pathogen attack. The adopted methods have been used for the alteration of plant immunity at several stages in different crops [163]. For example, wheat genotypes showed resistance to powdery mildew via the genome editing of the mildew resistance locus O (MLO) gene using the TALEN and CRISPR-Cas9 techniques [91]. Genome-editing technology has also been implicated in the generation of resistant plant lines against the bacterial leaf blight caused by Xanthomonas oryzae pv. oryzae [10,164]. The CRISPR/Cas9 system has been monitored for its ability to provide resistance against geminivirus infection, and geminivirus resistance has been established in both Arabidopsis and N. benthamiana by introducing sgRNA/Cas9 [165] (Table 5). Another successful modification was afforded by the promoter of CsLOB1, which rendered the resulting homozygous plants resistant to Citrus canker [157]. The comparative measured mutation rate was 3.2–3.9%, with no off-target effects. CRISPR-mediated editing in the grape cultivar “Chardonnay” showed that the targeted L-idonate dehydrogenase (IdnDH) gene exhibited a 100% mutation frequency rate
Recently, resistance to bacterial blight in rice has been improved through CRISPR/Cas9 by editing the SWEET11, SWEET13, and SWEET14 genes [166]. In tomatoes, downy mildew resistance 6 (DMR6) knockout mutants exhibited improved wide-spectrum resistance to multiple pathogens, including bacteria and oomycetes [167]. Similarly, the LATERAL ORGAN BOUNDARIES 1 transcription factor (CsLOB1) stimulates the proliferation of Xanthomonas citri ssp, which was reported as a causative agent of Citrus canker [168]. Moreover, several other studies have shown that CRISPR/Cas9 is effective in generating resistance against viruses in plants, such as the tomato yellow leaf curl virus (TYLCV) and the bean yellow dwarf virus (BeYDV) [165,169]. The delivery of sgRNAs targeting Cas9-pressing tobacco, including TYLCV, in the intergenic region, coat protein (CP), and the viral accumulation of several important viruses [169]. Furthermore, the ethylene-dependent pathway in rice has been modified and mutated successfully via CRISPR/Cas9 edition of the OsERF922 gene, with the resulting plants showing improved resistance to Magnaporthe oryzae [170].
Similarly, two genes of tobacco, NtPDS and NtPDR6, which encode pleiotropic drug resistance, were modified and mutated with CRISPR/Cas9, resulting in indel frequencies of 16.2–20.3% in protoplasts. Transgenic plants exhibited mutation rates of 81.8% and 87.5% in NtPDS and NtPDR6, respectively, whereas no significant effect was found near the off-target sites [104]. Furthermore, knockout of the gene encoding the eukaryotic translation initiation factor isoform 4E (eIF(iso)4E) in Arabidopsis using CRISPR/Cas9 resulted in the enhancement of turnip mosaic virus resistance but did not affect plant vigor [171]. A similar work reported in 2016 showed that the generation of two mutation sites in the eIF4E gene by CRISPR/Cas9 in cucumber resulted in resistance to the cucumber vein yellowing virus (CVYV), zucchini yellow mosaic virus (ZYMV), and papaya ring spot mosaic virus-W (PRSV-W) [172].
Such CRISPR/Cas9 applications collectively indicate that it is an essential tool of genome-editing technology and a key participant in the implementation of plant disease resistance.

3. Speed Breeding and MAS Using Genome-Editing Tools

The growing human population and changing environment entail several global concerns related to food security [184]. In the early 1990s, molecular markers were commonly used to select the most appropriate breeding lines [185], followed by genomics-assisted breeding in later years [186]. The CRISPR/Cpf1 tool was utilized for plant genome editing in 2016 [67]. Recently, marker-assisted selection (MAS) emerged as an important tool for genome editing. However, the method of implementation of this technique may vary with the advent of modern technologies [187,188]. A group of researchers presented rice as an example of how different mechanisms can be employed to develop an efficient tool to implement genetic variation for crop improvement [189]. Similarly, progress in maize breeding was made by integrating advances in sequencing, genotyping, and transformation, which includes doubled haploid technology and genome editing [190]. Recently, the “Sorghum QTL Atlas” provided an accessible research platform to deploy gene discovery among several species [191]. Subsequently, this concept was demonstrated in barley [192] and legume crops [193,194]. The availability of sequence information has led to more efficient breeding techniques.
“Speed breeding” (SB) is another magic tool that recently obtained notable attention. SB not only shortens the breeding cycle but also accelerates crop research through rapid generation advancement [195]. This technique utilizes artificial light coupled with temperature conditions to accelerate the crossing and inbreeding of various varieties. According to a recent study, the generation time is significantly reduced by providing a 22-h photoperiod and controlled temperature in several crops, such as spring bread wheat (Triticum aestivum), durum wheat (T. durum), barley (Hordeum vulgare), chickpea (Cicer arietinum), pea (Pisum sativum), canola (Brassica napus), the model grass Brachypodium distachyon, and the model legume Medicago truncatula, compared with the field or a greenhouse with no supplementary light [184,188,196]. Researchers believe that SB holds great potential via its integration with other modern crop technologies, such as high-throughput genotyping, genome editing, and genomic selection, to speed up the rate of crop improvement. The upcoming decade will witness the use of these powerful genome-editing technologies in combination with SB to enhance the speed and impact of better plant genotypes for farmers and consumers worldwide.

4. Speed Editing Strategy for Gene-Family Members

Recently, we developed a web tool to estimate the functional redundancy of rice genes, because more than 60% of the rice genome has multiple members in the same gene family, as assessed based on Pfam annotation. Functional redundancy associated with gene-family members is one of the main obstacles to crop improvement through gene-editing mediated by a loss-of-function method. Although a gene-editing system involving multiple genes was established in plant species, the editing of multiple genes at a time is generally a complex process. To achieve multiple-gene editing more effectively, we need to select candidate genes with functional redundancy more precisely by considering both protein sequence similarity and coexpression patterns among homologs. Although homologous genes account for more than half of genomes, 7075 out of 33,483 rice genes composing 2617 Pfam gene families retained a Pearson’s correlation coefficient (PCC) value of >0.7 for meta-expression data of anatomical samples, which suggests the existence of other genes with similar function; moreover, 8503 genes exhibited a very low level of expression, which hampers the estimation of their function based on expression data. Therefore, 46.5% of rice genes with gene-family members might not be suitable targets for gene-editing applications using a single target. In addition, the 7075 genes with other family members in the genome having higher similarity in both sequence and expression patterns are more probable targets for multiple-gene editing (Figure 6) [197]. For example, OsMADS63 genes, which share expression in mature pollen with OsMADS62, did not yield a defect in pollen development, whereas multiple mutations of OsMADS62 and OsMADS63 with a PCC of 0.977 caused a severe defect in late pollen development, and RUPO mutation (PCC, 0.335) over LOC_Os03g55210 in the same family led to a severe defect in late pollen development and did not require multiple-gene editing. Therefore, accurate estimation of gene redundancy within a family will accelerate crop improvement through gene-editing systems. This web tool (CAFRI-Rice, http://cafri-rice.khu.ac.kr/) is only available for rice, but we expect its expansion to other crop species [197].

5. Future Directions

The adopted CRISPR system and its usage will promote the rapid progress of crop breeding and functional genomics. Recently, new and versatile breeding technologies have been implemented to facilitate the engineering of multiple genetic loci in different breeding varieties, which will improve food security and strengthen crop amelioration. Moreover, the perusal of the literature for genomic sequences and their functions is a prerequisite for efficient genome editing. In the future, we will likely witness the increased use of CRISPR for clarifying genomic structures and their role in plants, such as the transcriptional regulation of Cas9 and Cpf1, the monitoring of genetic loci and mechanisms, and the regulation of promoter activity. Moreover, it will also include the modification and identification of epigenetic behavior in communicating the stable relationships between single-nucleotide polymorphisms (SNPs), which are investigated by genetic traits, and genome-wide association studies. Interestingly, this technique was designed to achieve phenotypic characterization in the T0 generation by engineering a genome-wide mutant library in rice. Upon consideration of the highly efficient editing achieved in the T0 generation, CRISPR/Cas9 was used to engineer a genome-wide mutant library in rapeseed, which will promote gene characterization and its beneficial applications at a later stage.
The CRISPR technology can classify any new crop traits in the category of plant synthetic biology (Figure 7). In our opinion, the saturation mutagenesis induced by CRISPR could be used to develop any desired plant protein when a proper selection tool is available. The use of this “faster and cheaper” method of evolution to optimize the role of metabolic enzymes in traits, such as crop production, quality, and disease resistance, should accelerate crop development; however, the CRISPR-associated technology would need to be strengthened. For example, improvements in the transformation methods and delivery of CRISPR/Cas agents to target cells will enable CRISPR in different tissues, including germline cells, and will increase the compatibility of plant species. However, the off-target issue is a big challenge in the application of gene-editing technology and a recent whole-genome sequencing analysis of CRISPR/Cas9-edited cotton plants revealed rare off-target mutations [198]. The detailed strategy to increase on-target and reduce off-target effects of CRISPR/Cas9 was recently well reviewed [199]. Targeted genome editing in rice using chemically modified donor DNA, which are designed for UTR or prompter region and the homology-directed repair method, was successful [200] and further improvements are expected in the future. Creating a large population of CRISPR/Cas9-driven mutagenesis of promoters for developmental genes of tomato contributes to increased genetic variations [201]. It is reasonable to expect that the development of various precision genome-editing technologies for targeted and precise gene/allele replacement, in combination with conventional breeding practices, will expedite the breeding of diverse elite crop varieties for the development of sustainable agriculture.

6. Conclusions

Genome editing is becoming the most used and versatile tool for crop improvement and functional genomics. The attractive survival landscapes, such as the efficiency, multiplexing, integrity, and simplicity, as well as the highly specific nature, of the genome-editing technologies mentioned here indicate the manner in which crop breeding is carried out and pave the way for plant breeding for the next generations. This new strategy for crop improvement has proven to be efficacious based on a review of the literature on transcriptomics, biotechnology, genomics, and phonemics. The regulation of transgenic crops was also coherently simplified to support the rapid progression of this technology and render these crops acceptable for consumer usage. In addition to these social and technical challenges, the CRISPR technology was used for the first time to edit plant genomes. Therefore, the use of genome editing on a large scale for crop improvement is already a reality. The journey of genome editing raises ethical questions that need to be addressed by researchers and society on a massive scale.

Author Contributions

S.A. and K.-H.J. Designed the study; S.A., S.S. (Saba Saleem), A.F., A.M., S.U.K., M.H.U.K. and F.M.-P. wrote the MS; S.U.K., M.H.U.K., S.S. (Sumbul Saeed), M.K., & M.R., help in diagrams. S.A., F.M.-P., W.-J.H., and K.-H.J., M.R., and M.K., revised the MS. Supervision K.-H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This Research was funded by the National Research Foundation of Korea (2018R1A4A1025158), the Next-Generation BioGreen 21 Program (PJ01325901 and PJ01366401), and a Program from the Rural Development Administration (PJ01492703).

Acknowledgments

The authors are thankful to an anonymous reviewer for their comments and critical reading of the manuscript. There is no specific funding for this article.

Conflicts of Interest

All authors declare that there is no conflict of interest.

Abbreviations

ZFNZinc-Finger Nucleases
TALENsTranscriptional activator-like Effector Nucleases
CRISPRClustered Regularly Interspaced Short Palindromic Repeats (CRISPR)
Cas9CRISPR-associated Proteins
Cpf1CRISPR-associated endonuclease in Prevotella and Francisella
DSBDouble-Strand Breaks
NHEJNonhomologous end jointing
HDRHomology-directed repair mechanism
GMOGenetically Modified Organism
SBSpeed breeding
RVDRepeat variable di-residue

References

  1. Ray, D.K.; Mueller, N.D.; West, P.C.; Foley, J.A. Yield Trends Are Insufficient to Double Global Crop Production by 2050. PLoS ONE 2013, 8, e66428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Röös, E.; Bajželj, B.; Smith, P.; Patel, M.; Little, D.; Garnett, T. Greedy or needy? Land use and climate impacts of food in 2050 under different livestock futures. Glob. Environ. Chang. 2017, 47, 1–2. [Google Scholar] [CrossRef]
  3. Zhang, H.; Zhang, J.; Lang, Z.; Botella, J.R.; Zhu, J.K. Genome Editing—Principles and Applications for Functional Genomics Research and Crop Improvement. CRC. Crit. Rev. Plant Sci. 2017, 36, 291–309. [Google Scholar] [CrossRef]
  4. Gaj, T.; Gersbach, C.A.; Barbas, C.F. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rai, K.M.; Ghose, K.; Rai, A.; Singh, H.; Srivastava, R.; Mendu, V. Genome engineering tools in plant synthetic biology. In Current Developments in Biotechnology and Bioengineering: Synthetic Biology. Cell Eng. Bioprocess. Technol. 2018. [Google Scholar] [CrossRef]
  6. Osakabe, K.; Osakabe, Y.; Toki, S. Site-directed mutagenesis in Arabidopsis using custom-designed zinc finger nucleases. Proc. Natl. Acad. Sci. USA 2010, 107, 12034–12039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Petolino, J.F.; Worden, A.; Curlee, K.; Connell, J.; Moynahan, T.L.S.; Larsen, C.; Russell, S. Zinc finger nuclease-mediated transgene deletion. Plant Mol. Biol. 2010, 73, 617–628. [Google Scholar] [CrossRef]
  8. 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]
  9. Boch, J.; Scholze, H.; Schornack, S.; Landgraf, A.; Hahn, S.; Kay, S.; Lahaye, T.; Nickstadt, A.; Bonas, U. Breaking the code of DNA binding specificity of TAL-type III effectors. Science 2009, 326, 1509–1512. [Google Scholar] [CrossRef]
  10. 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. [Google Scholar] [CrossRef]
  11. Kopischke, S.; Schüßler, E.; Althoff, F.; Zachgo, S. TALEN-mediated genome-editing approaches in the liverwort Marchantia polymorpha yield high efficiencies for targeted mutagenesis. Plant Methods 2017, 13, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bortesi, L.; Fischer, R. The CRISPR/Cas9 system for plant genome editing and beyond. Biotechnol. Adv. 2015, 33, 41–52. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Y.; Pribil, M.; Palmgren, M.; Gao, C. A CRISPR way for accelerating improvement of food crops. Nat. Food 2020, 1, 200–205. [Google Scholar] [CrossRef]
  14. Butt, H.; Zaidi, S.S.A.; Hassan, N.; Mahfouz, M. CRISPR-Based Directed Evolution for Crop Improvement. Trends Biotechnol. 2020, 38, 236–240. [Google Scholar] [CrossRef] [PubMed]
  15. Puchta, H. Applying CRISPR/Cas for genome engineering in plants: The best is yet to come. Curr. Opin. Plant Biol. 2017, 36, 1–8. [Google Scholar] [CrossRef] [PubMed]
  16. Ahmar, S.; Gill, R.A.; Jung, K.H.; Faheem, A.; Qasim, M.U.; Mubeen, M.; Zhou, W. Conventional and molecular techniques from simple breeding to speed breeding in crop plants: Recent advances and future outlook. Int. J. Mol. Sci. 2020, 21, 2590. [Google Scholar] [CrossRef] [Green Version]
  17. Mishra, R.; Zhao, K. Genome editing technologies and their applications in crop improvement. Plant Biotechnol. Rep. 2018, 12, 57–68. [Google Scholar] [CrossRef]
  18. 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] [Green Version]
  19. Zhang, F.; Maeder, M.L.; Unger-Wallaced, E.; Hoshaw, J.P.; Reyon, D.; Christian, M.; Li, X.; Pierick, C.J.; Dobbs, D.; Peterson, T.; et al. High frequency targeted mutagenesis in Arabidopsis thaliana using zinc finger nucleases. Proc. Natl. Acad. Sci. USA 2010, 107, 12028–12033. [Google Scholar] [CrossRef] [Green Version]
  20. Puchta, H. The repair of double-strand breaks in plants: Mechanisms and consequences for genome evolution. J. Exp. Bot. 2005, 56, 1–4. [Google Scholar] [CrossRef] [Green Version]
  21. Durai, S.; Mani, M.; Kandavelou, K.; Wu, J.; Porteus, M.H.; Chandrasegaran, S. Zinc finger nucleases: Custom-designed molecular scissors for genome engineering of plant and mammalian cells. Nucleic Acids Res. 2005, 33, 5978–5990. [Google Scholar] [CrossRef] [PubMed]
  22. Papworth, M.; Kolasinska, P.; Minczuk, M. Designer zinc-finger proteins and their applications. Gene 2006, 366, 27–38. [Google Scholar] [CrossRef] [PubMed]
  23. Carlson, D.F.; Fahrenkrug, S.C.; Hackett, P.B. Targeting DNA with fingers and TALENs. Mol. Ther. Nucleic Acids 2012. [Google Scholar] [CrossRef] [PubMed]
  24. Carroll, D.; Morton, J.J.; Beumer, K.J.; Segal, D.J. Design, construction and in vitro testing of zinc finger nucleases. Nat. Protoc. 2006, 1, 1329–1341. [Google Scholar] [CrossRef]
  25. Minczuk, M.; Papworth, M.A.; Miller, J.C.; Murphy, M.P.; Klug, A. Development of a single-chain, quasi-dimeric zinc-finger nuclease for the selective degradation of mutated human mitochondrial DNA. Nucleic Acids Res. 2008, 36, 3926–3938. [Google Scholar] [CrossRef] [Green Version]
  26. Gaj, T.; Guo, J.; Kato, Y.; Sirk, S.J.; Barbas, C.F. Targeted gene knockout by direct delivery of zinc-finger nuclease proteins. Nat. Methods 2012, 9, 805–807. [Google Scholar] [CrossRef] [Green Version]
  27. Bedell, V.M.; Wang, Y.; Campbell, J.M.; Poshusta, T.L.; Starker, C.G.; Krug, R.G.; Tan, W.; Penheiter, S.G.; Ma, A.C.; Leung, A.Y.H.; et al. In vivo genome editing using a high-efficiency TALEN system. Nature 2012, 491, 114–118. [Google Scholar] [CrossRef] [Green Version]
  28. 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] [Green Version]
  29. 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]
  30. Römer, P.; Hahn, S.; Jordan, T.; Strauß, T.; Bonas, U.; Lahaye, T. Plant pathogen recognition mediated by promoter activation of the pepper Bs3 resistance gene. Science 2007, 318, 645–648. [Google Scholar] [CrossRef] [Green Version]
  31. Lamb, B.M.; Mercer, A.C.; Barbas III, C.F. Directed evolution of the TALE N-terminal domain for recognition of all 5′ bases. Nucleic Acids Res. 2013, 41, 9779–9785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Gasiunas, G.; Barrangou, R.; Horvath, P.; Siksnys, V. Cas9-crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proc. Natl. Acad. Sci. USA 2012, 109, E2579–E2586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chylinski, K.; Makarova, K.S.; Charpentier, E.; Koonin, E.V. Classification and evolution of type II CRISPR-Cas systems. Nucleic Acids Res. 2014, 42, 6091–6105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Smirnov, A.V.; Yunusova, A.M.; Lukyanchikova, V.A.; Battulin, N.R. CRISPR/Cas9, a universal tool for genomic engineering. Russ. J. Genet. Appl. Res. 2017, 7, 440–458. [Google Scholar] [CrossRef]
  35. Cheng, A.W.; Wang, H.; Yang, H.; Shi, L.; Katz, Y.; Theunissen, T.W.; Rangarajan, S.; Shivalila, C.S.; Dadon, D.B.; Jaenisch, R. Multiplexed activation of endogenous genes by CRISPR-on, an RNA-guided transcriptional activator system. Cell Res. 2013, 23, 1163–1171. [Google Scholar] [CrossRef] [PubMed]
  36. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef] [Green Version]
  37. Makarova, K.S.; Haft, D.H.; Barrangou, R.; Brouns, S.J.J.; Charpentier, E.; Horvath, P.; Moineau, S.; Mojica, F.J.M.; Wolf, Y.I.; Yakunin, A.F.; et al. Evolution and classification of the CRISPR-Cas systems. Nat. Rev. Microbiol. 2011, 9, 467–477. [Google Scholar] [CrossRef] [Green Version]
  38. Makarova, K.S.; Grishin, N.V.; Shabalina, S.A.; Wolf, Y.I.; Koonin, E.V. A putative RNA-interference-based immune system in prokaryotes: Computational analysis of the predicted enzymatic machinery, functional analogies with eukaryotic RNAi, and hypothetical mechanisms of action. Biol. Direct 2006, 1, 440. [Google Scholar] [CrossRef] [Green Version]
  39. Lemmon, Z.H.; Reem, N.T.; Dalrymple, J.; Soyk, S.; Swartwood, K.E.; Rodriguez-Leal, D.; Van Eck, J.; Lippman, Z.B. Rapid improvement of domestication traits in an orphan crop by genome editing. Nat. Plants 2018, 4, 766–770. [Google Scholar] [CrossRef]
  40. Symington, L.S.; Gautier, J. Double-strand break end resection and repair pathway choice. Annu. Rev. Genet. 2011, 45, 247–271. [Google Scholar] [CrossRef]
  41. Zha, S.; Boboila, C.; Alt, F.W. Mre11: Roles in DNA repair beyond homologous recombination. Nat. Struct. Mol. Biol. 2009, 16, 798–800. [Google Scholar] [CrossRef] [PubMed]
  42. 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]
  43. Hilscher, J.; Bürstmayr, H.; Stoger, E. Targeted modification of plant genomes for precision crop breeding. Biotechnol. J. 2017, 12, 1600173. [Google Scholar] [CrossRef]
  44. Ma, X.; Chen, X.; Jin, Y.; Ge, W.; Wang, W.; Kong, L.; Ji, J.; Guo, X.; Huang, J.; Feng, X.H.; et al. Small molecules promote CRISPR-Cpf1-mediated genome editing in human pluripotent stem cells. Nat. Commun. 2018, 9, 1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Riesenberg, S.; Maricic, T. Targeting repair pathways with small molecules increases precise genome editing in pluripotent stem cells. Nat. Commun. 2018, 9, 2164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Safari, F.; Zare, K.; Negahdaripour, M.; Barekati-Mowahed, M.; Ghasemi, Y. CRISPR Cpf1 proteins: Structure, function and implications for genome editing. Cell Biosci. 2019, 9, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Malzahn, A.A.; Tang, X.; Lee, K.; Ren, Q.; Sretenovic, S.; Zhang, Y.; Chen, H.; Kang, M.; Bao, Y.; Zheng, X.; et al. Application of CRISPR-Cas12a temperature sensitivity for improved genome editing in rice, maize, and Arabidopsis. BMC Biol. 2019, 17, 9. [Google Scholar] [CrossRef] [Green Version]
  48. Bernabé-Orts, J.M.; Casas-Rodrigo, I.; Minguet, E.G.; Landolfi, V.; Garcia-Carpintero, V.; Gianoglio, S.; Vázquez-Vilar, M.; Granell, A.; Orzaez, D. Assessment of Cas12a-mediated gene editing efficiency in plants. Plant Biotechnol. J. 2019, 17, 1971–1984. [Google Scholar] [CrossRef] [Green Version]
  49. Bayat, H.; Modarressi, M.H.; Rahimpour, A. The Conspicuity of CRISPR-Cpf1 System as a Significant Breakthrough in Genome Editing. Curr. Microbiol. 2018, 75, 107–115. [Google Scholar] [CrossRef]
  50. Ding, D.; Chen, K.; Chen, Y.; Li, H.; Xie, K. Engineering Introns to Express RNA Guides for Cas9- and Cpf1-Mediated Multiplex Genome Editing. Mol. Plant 2018, 11, 542–552. [Google Scholar] [CrossRef] [Green Version]
  51. Li, S.; Zhang, X.; Wang, W.; Guo, X.; Wu, Z.; Du, W.; Zhao, Y.; Xia, L. Expanding the scope of CRISPR/Cpf1-mediated genome editing in rice. Mol. Plant 2018, 11, 995–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Li, T.; Zhu, L.; Xiao, B.; Gong, Z.; Liao, Q.; Guo, J. CRISPR-Cpf1-mediated genome editing and gene regulation in human cells. Biotechnol. Adv. 2018, 37, 21–27. [Google Scholar] [CrossRef] [PubMed]
  53. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Publisher Correction: Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 2018, 559, E8. [Google Scholar] [CrossRef]
  54. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef]
  56. 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]
  57. Upadhyay, S.K.; Kumar, J.; Alok, A.; Tuli, R. RNA-Guided genome editing for target gene mutations in wheat. G3 Genes Genomes Genet. 2013, 3, 2233–2238. [Google Scholar] [CrossRef] [Green Version]
  58. Eş, I.; Gavahian, M.; Marti-Quijal, F.J.; Lorenzo, J.M.; Mousavi Khaneghah, A.; 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]
  59. Sauer, N.J.; Mozoruk, J.; Miller, R.B.; Warburg, Z.J.; Walker, K.A.; Beetham, P.R.; Schöpke, C.R.; Gocal, G.F.W. Oligonucleotide-directed mutagenesis for precision gene editing. Plant Biotechnol. J. 2016, 14, 496–502. [Google Scholar] [CrossRef] [Green Version]
  60. Kumar, V.; Jain, M. The CRISPR-Cas system for plant genome editing: Advances and opportunities. J. Exp. Bot. 2015, 66, 47–57. [Google Scholar] [CrossRef] [Green Version]
  61. Hsu, P.D.; Scott, D.A.; Weinstein, J.A.; Ran, F.A.; Konermann, S.; Agarwala, V.; Li, Y.; Fine, E.J.; Wu, X.; Shalem, O.; et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 2013, 31, 827–832. [Google Scholar] [CrossRef]
  62. 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]
  63. Swarts, D.C.; Jinek, M. Cas9 versus Cas12a/Cpf1: Structure–function comparisons and implications for genome editing. Wiley Interdiscip. Rev. RNA 2018, 9, e1481. [Google Scholar] [CrossRef] [PubMed]
  64. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  65. Chen, B.; Hu, J.; Almeida, R.; Liu, H.; Balakrishnan, S.; Covill-Cooke, C.; Lim, W.A.; Huang, B. Expanding the CRISPR imaging toolset with Staphylococcus aureus Cas9 for simultaneous imaging of multiple genomic loci. Nucleic Acids Res. 2016, 44, e75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. 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]
  67. Endo, A.; Masafumi, M.; Kaya, H.; Toki, S. Efficient targeted mutagenesis of rice and tobacco genomes using Cpf1 from Francisella novicida. Sci. Rep. 2016, 6, 38169. [Google Scholar] [CrossRef] [Green Version]
  68. Khatodia, S.; Bhatotia, K.; Passricha, N.; Khurana, S.M.P.; Tuteja, N. The CRISPR/Cas genome-editing tool: Application in improvement of crops. Front. Plant Sci. 2016, 7, 506. [Google Scholar] [CrossRef] [Green Version]
  69. Khandagale, K.; Nadaf, A. Genome editing for targeted improvement of plants. Plant Biotechnol. Rep. 2016, 10, 327–343. [Google Scholar] [CrossRef]
  70. Song, G.; Jia, M.; Chen, K.; Kong, X.; Khattak, B.; Xie, C.; Li, A.; Mao, L. CRISPR/Cas9: A powerful tool for crop genome editing. Crop J. 2016, 4, 75–82. [Google Scholar] [CrossRef] [Green Version]
  71. 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]
  72. Lloyd, A.; Plaisier, C.L.; Carroll, D.; Drews, G.N. Targeted mutagenesis using zinc-finger nucleases in Arabidopsis. Proc. Natl. Acad. Sci. USA 2005, 102, 2232–2237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Wright, D.A.; Townsend, J.A.; Winfrey, R.J.; 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]
  74. 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] [PubMed]
  75. De Pater, S.; Pinas, J.E.; Hooykaas, P.J.J.; van der Zaal, B.J. ZFN-mediated gene targeting of the Arabidopsis protoporphyrinogen oxidase gene through Agrobacterium-mediated floral dip transformation. Plant Biotechnol. J. 2013, 11, 510–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Carroll, D. Progress and prospects: Zinc-finger nucleases as gene therapy agents. Gene Ther. 2008, 15, 1463–1468. [Google Scholar] [CrossRef] [Green Version]
  77. 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] [Green Version]
  78. Fu, Y.; Xiao, M.; Yu, H.; Mason, A.S.; Yin, J.; Li, J.; Zhang, D.; Fu, D. Small RNA changes in synthetic Brassica napus. Planta 2016, 244, 607–622. [Google Scholar] [CrossRef]
  79. Hegedus, D.; Yu, M.; Baldwin, D.; Gruber, M.; Sharpe, A.; Parkin, I.; Whitwill, S.; Lydiate, D. Molecular characterization of Brassica napus NAC domain transcriptional activators induced in response to biotic and abiotic stress. Plant Mol. Biol. 2003, 53, 383–397. [Google Scholar] [CrossRef]
  80. Gupta, M.; Dekelver, R.C.; Palta, A.; Clifford, C.; Gopalan, S.; Miller, J.C.; Novak, S.; Desloover, D.; Gachotte, D.; Connell, J.; et al. Transcriptional activation of Brassica napusβ-ketoacyl-ACP synthase II with an engineered zinc finger protein transcription factor. Plant Biotechnol. J. 2012, 10, 783–791. [Google Scholar] [CrossRef]
  81. Christian, M.; Cermak, T.; Doyle, E.L.; Schmidt, C.; Zhang, F.; Hummel, A.; Bogdanove, A.J.; Voytas, D.F. Targeting DNA double-strand breaks with TAL effector nucleases. Genetics 2010, 186, 757–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Belhaj, K.; Chaparro-Garcia, A.; Kamoun, S.; Nekrasov, V. Plant genome editing made easy: Targeted mutagenesis in model and crop plants using the CRISPR/Cas system. Plant Methods 2013, 9, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Christian, M.; Cermak, T.; Doyle, E.L.; Schmidt, C.; Zhang, F.; Hummel, A.; Bogdanove, A.J.; Voytas, D.F. TAL Effector Nucleases Create Targeted DNA Double-strand Breaks. Genetics 2010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Mussolino, C.; Morbitzer, R.; Lütge, F.; Dannemann, N.; Lahaye, T.; Cathomen, T. A novel TALE nuclease scaffold enables high genome editing activity in combination with low toxicity. Nucleic Acids Res. 2011, 39, 9283–9293. [Google Scholar] [CrossRef] [PubMed]
  85. Miller, J.C.; Tan, S.; Qiao, G.; Barlow, K.A.; Wang, J.; Xia, D.F.; Meng, X.; Paschon, D.E.; Leung, E.; Hinkley, S.J.; et al. A TALE nuclease architecture for efficient genome editing. Nat. Biotechnol. 2011, 29, 143–148. [Google Scholar] [CrossRef] [PubMed]
  86. Chaudhary, K.; Pratap, D.; Sharma, P.K. Transcription activator-like effector nucleases (TALENs): An efficient tool for plant genome editing. Eng. Life Sci. 2016, 16, 330–337. [Google Scholar] [CrossRef] [Green Version]
  87. Lor, V.S.; Starker, C.G.; Voytas, D.F.; Weiss, D.; Olszewski, N.E. Targeted mutagenesis of the tomato procera gene using transcription activator-like effector nucleases. Plant Physiol. 2014, 166, 1288–1291. [Google Scholar] [CrossRef] [Green Version]
  88. 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]
  89. 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]
  90. 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]
  91. 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] [PubMed]
  92. Butler, N.M.; Baltes, N.J.; Voytas, D.F.; Douches, D.S. Geminivirus-mediated genome editing in potato (Solanum tuberosum l.) using sequence-specific nucleases. Front. Plant Sci. 2016, 7, 1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. 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] [PubMed]
  94. Jung, J.H.; Altpeter, F. TALEN mediated targeted mutagenesis of the caffeic acid O-methyltransferase in highly polyploid sugarcane improves cell wall composition for production of bioethanol. Plant Mol. Biol. 2016, 92, 131–142. [Google Scholar] [CrossRef] [Green Version]
  95. Shan, Q.; Zhang, Y.; Chen, K.; Zhang, K.; Gao, C. Creation of fragrant rice by targeted knockout of the OsBADH2 gene using TALEN technology. Plant Biotechnol. J. 2015, 13, 791–800. [Google Scholar] [CrossRef]
  96. 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]
  97. Jiang, W.Z.; Henry, I.M.; Lynagh, P.G.; Comai, L.; Cahoon, E.B.; Weeks, D.P. Significant enhancement of fatty acid composition in seeds of the allohexaploid, Camelina sativa, using CRISPR/Cas9 gene editing. Plant Biotechnol. J. 2017, 15, 648–657. [Google Scholar] [CrossRef] [Green Version]
  98. Aznar-Moreno, J.A.; Durrett, T.P. Simultaneous Targeting of Multiple Gene Homeologs to Alter Seed Oil Production in Camelina sativa. Plant Cell Physiol. 2017, 58, 1260–1267. [Google Scholar] [CrossRef]
  99. Fan, D.; Liu, T.; Li, C.; Jiao, B.; Li, S.; Hou, Y.; Luo, K. Efficient CRISPR/Cas9-mediated Targeted Mutagenesis in Populus in the First Generation. Sci. Rep. 2015, 5, 12217. [Google Scholar] [CrossRef]
  100. Hyun, Y.; Kim, J.; Cho, S.W.; Choi, Y.; Kim, J.S.; Coupland, G. Site-directed mutagenesis in Arabidopsis thaliana using dividing tissue-targeted RGEN of the CRISPR/Cas system to generate heritable null alleles. Planta 2014, 241, 271–284. [Google Scholar] [CrossRef] [Green Version]
  101. Ellison, E.E.; Nagalakshmi, U.; Gamo, M.E.; Huang, P.; Dinesh-Kumar, S.; Voytas, D.F. Multiplexed heritable gene editing using RNA viruses and mobile single guide RNAs. Nat. Plants 2020, 6, 620–624. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, Z.; Mao, Y.; Ha, S.; Liu, W.; Botella, J.R.; Zhu, J.K. A multiplex CRISPR/Cas9 platform for fast and efficient editing of multiple genes in Arabidopsis. Plant Cell Rep. 2016, 35, 1519–1533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Nekrasov, V.; Staskawicz, B.; Weigel, D.; Jones, J.D.G.; Kamoun, S. Targeted mutagenesis in the model plant Nicotiana benthamiana using Cas9 RNA-guided endonuclease. Nat. Biotechnol. 2013, 31, 691–693. [Google Scholar] [CrossRef] [PubMed]
  104. Gao, J.; Wang, G.; Ma, S.; Xie, X.; Wu, X.; Zhang, X.; Wu, Y.; Zhao, P.; Xia, Q. CRISPR/Cas9-mediated targeted mutagenesis in Nicotiana tabacum. Plant Mol. Biol. 2015, 87, 99–110. [Google Scholar] [CrossRef] [PubMed]
  105. 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]
  106. Zhou, H.; Liu, B.; Weeks, D.P.; Spalding, M.H.; Yang, B. Large chromosomal deletions and heritable small genetic changes induced by CRISPR/Cas9 in rice. Nucleic Acids Res. 2014, 42, 10903–10914. [Google Scholar] [CrossRef]
  107. Xu, R.F.; Li, H.; Qin, R.Y.; Li, J.; Qiu, C.H.; Yang, Y.C.; Ma, H.; Li, L.; Wei, P.C.; Yang, J.B. Generation of inheritable and “transgene clean” targeted genome-modified rice in later generations using the CRISPR/Cas9 system. Sci. Rep. 2015, 19, 11491. [Google Scholar] [CrossRef] [Green Version]
  108. Lawrenson, T.; Shorinola, O.; Stacey, N.; Li, C.; Østergaard, L.; Patron, N.; Uauy, C.; Harwood, W. Induction of targeted, heritable mutations in barley and Brassica oleracea using RNA-guided Cas9 nuclease. Genome Biol. 2015, 16, 258. [Google Scholar] [CrossRef] [Green Version]
  109. Morineau, C.; Bellec, Y.; Tellier, F.; Gissot, L.; Kelemen, Z.; Nogué, F.; Faure, J.D. Selective gene dosage by CRISPR-Cas9 genome editing in hexaploid Camelina sativa. Plant Biotechnol. J. 2017, 15, 729–739. [Google Scholar] [CrossRef] [Green Version]
  110. Okuzaki, A.; Ogawa, T.; Koizuka, C.; Kaneko, K.; Inaba, M.; Imamura, J.; Koizuka, N. CRISPR/Cas9-mediated genome editing of the fatty acid desaturase 2 gene in Brassica napus. Plant Physiol. Biochem. 2018, 131, 63–69. [Google Scholar] [CrossRef]
  111. Long, S.P.; Marshall-Colon, A.; Zhu, X.G. Meeting the global food demand of the future by engineering crop photosynthesis and yield potential. Cell 2015, 161, 56–66. [Google Scholar] [CrossRef] [Green Version]
  112. Shen, L.; Wang, C.; Fu, Y.; Wang, J.; Liu, Q.; Zhang, X.; Yan, C.; Qian, Q.; Wang, K. QTL editing confers opposing yield performance in different rice varieties. J. Integr. Plant Biol. 2018, 60, 89–93. [Google Scholar] [CrossRef] [PubMed]
  113. Li, R.; Li, R.; Li, X.; Fu, D.; Zhu, B.; Tian, H.; Luo, Y.; Zhu, H. Multiplexed CRISPR/Cas9-mediated metabolic engineering of γ-aminobutyric acid levels in Solanum lycopersicum. Plant Biotechnol. J. 2018, 16, 415–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. 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]
  115. Ueta, R.; Abe, C.; Watanabe, T.; Sugano, S.S.; Ishihara, R.; Ezura, H.; Osakabe, Y.; Osakabe, K. Rapid breeding of parthenocarpic tomato plants using CRISPR/Cas9. Sci. Rep. 2017, 7, 507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Soyk, S.; Müller, N.A.; Park, S.J.; Schmalenbach, I.; Jiang, K.; Hayama, R.; Zhang, L.; Van Eck, J.; Jiménez-Gómez, J.M.; Lippman, Z.B. Variation in the flowering gene SELF PRUNING 5G promotes day-neutrality and early yield in tomato. Nat. Genet. 2017, 49, 162–168. [Google Scholar] [CrossRef] [PubMed]
  117. Brooks, C.; Nekrasov, V.; Lippman, Z.B.; Van Eck, J. Efficient gene editing in tomato in the first generation using the clustered regularly interspaced short palindromic repeats/CRISPR-associated9 system. Plant Physiol. 2014, 166, 1292–1297. [Google Scholar] [CrossRef] [Green Version]
  118. 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]
  119. Jia, H.; Nian, W. Targeted genome editing of sweet orange using Cas9/sgRNA. PLoS ONE 2014, 9, e93806. [Google Scholar] [CrossRef] [Green Version]
  120. Malnoy, M.; Viola, R.; Jung, M.H.; Koo, O.J.; Kim, S.; Kim, J.S.; Velasco, R.; Kanchiswamy, C.N. DNA-free genetically edited grapevine and apple protoplast using CRISPR/Cas9 ribonucleoproteins. Front. Plant Sci. 2016, 7, 1904. [Google Scholar] [CrossRef]
  121. Toda, E.; Koiso, N.; Takebayashi, A.; Ichikawa, M.; Kiba, T.; Osakabe, K.; Osakabe, Y.; Sakakibara, H.; Kato, N.; Okamoto, T. An efficient DNA- and selectable-marker-free genome-editing system using zygotes in rice. Nat. Plants 2019, 5, 363. [Google Scholar] [CrossRef] [PubMed]
  122. Espley, R.V.; Leif, D.; Plunkett, B.; McGhie, T.; Henry-Kirk, R.; Hall, M.; Johnston, J.W.; Punter, M.P.; Boldingh, H.; Nardozza, S.; et al. Red to Brown: An Elevated Anthocyanic Response in Apple Drives Ethylene to Advance Maturity and Fruit Flesh Browning. Front. Plant Sci. 2019, 10, 1248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. 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] [PubMed]
  124. Begemann, M.B.; Gray, B.N.; January, E.; Gordon, G.C.; He, Y.; Liu, H.; Wu, X.; Brutnell, T.P.; Mockler, T.C.; Oufattole, M. Precise insertion and guided editing of higher plant genomes using Cpf1 CRISPR nucleases. Sci. Rep. 2017, 7, 11606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Wang, M.; Mao, Y.; Lu, Y.; Wang, Z.; Tao, X.; Zhu, J.K. Multiplex gene editing in rice with simplified CRISPR-Cpf1 and CRISPR-Cas9 systems. J. Integr. Plant Biol. 2018, 60, 626–631. [Google Scholar] [CrossRef]
  126. Naim, F.; Dugdale, B.; Kleidon, J.; Brinin, A.; Shand, K.; Waterhouse, P.; Dale, J. Gene editing the phytoene desaturase alleles of Cavendish banana using CRISPR/Cas9. Transgenic Res. 2018, 27, 451–460. [Google Scholar] [CrossRef] [Green Version]
  127. Kim, H.; Kim, S.T.; Ryu, J.; Kang, B.C.; Kim, J.S.; Kim, S.G. CRISPR/Cpf1-mediated DNA-free plant genome editing. Nat. Commun. 2017, 8, 14406. [Google Scholar] [CrossRef] [Green Version]
  128. Xu, R.; Qin, R.; Li, H.; Li, D.; Li, L.; Wei, P.; Yang, J. Generation of targeted mutant rice using a CRISPR-Cpf1 system. Plant Biotechnol. J. 2017, 15, 713–717. [Google Scholar] [CrossRef] [Green Version]
  129. Jiang, W.; Zhou, H.; Bi, H.; Fromm, M.; Yang, B.; Weeks, D.P. Demonstration of CRISPR/Cas9/sgRNA-mediated targeted gene modification in Arabidopsis, tobacco, sorghum and rice. Nucleic Acids Res. 2013, 41, e188. [Google Scholar] [CrossRef]
  130. Li, M.; Li, X.; Zhou, Z.; Wu, P.; Fang, M.; Pan, X.; Lin, Q.; Luo, W.; Wu, G.; Li, H. Reassessment of the four yield-related genes Gn1a, DEP1, GS3, and IPA1 in rice using a CRISPR/Cas9 system. Front. Plant Sci. 2016, 7, 377. [Google Scholar] [CrossRef] [Green Version]
  131. Wang, J.; Cappa, J.J.; Harris, J.P.; Edger, P.P.; Zhou, W.; Pires, J.C.; Adair, M.; Unruh, S.A.; Simmons, M.P.; Schiavon, M.; et al. Transcriptome-wide comparison of selenium hyperaccumulator and nonaccumulator Stanleya species provides new insight into key processes mediating the hyperaccumulation syndrome. Plant Biotechnol. J. 2018, 16, 1582–1594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Cai, Y.; Chen, L.; Liu, X.; Guo, C.; Sun, S.; Wu, C.; Jiang, B.; Han, T.; Hou, W. CRISPR/Cas9-mediated targeted mutagenesis of GmFT2a delays flowering time in soya bean. Plant Biotechnol. J. 2018, 16, 176–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Ito, Y.; Nishizawa-Yokoi, A.; Endo, M.; Mikami, M.; Toki, S. CRISPR/Cas9-mediated mutagenesis of the RIN locus that regulates tomato fruit ripening. Biochem. Biophys. Res. Commun. 2015, 467, 76–82. [Google Scholar] [CrossRef] [PubMed]
  134. Sander, J.D.; Joung, J.K. CRISPR-Cas systems for editing, regulating and targeting genomes. Nat. Biotechnol. 2014, 32, 347–355. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, J.; Zhang, H.; Botella, J.R.; Zhu, J.K. Generation of new glutinous rice by CRISPR/Cas9-targeted mutagenesis of the Waxy gene in elite rice varieties. J. Integr. Plant Biol. 2018, 60, 369–375. [Google Scholar] [CrossRef]
  136. Li, X.; Zhou, W.; Ren, Y.; Tian, X.; Lv, T.; Wang, Z.; Fang, J.; Chu, C.; Yang, J.; Bu, Q. High-efficiency breeding of early-maturing rice cultivars via CRISPR/Cas9-mediated genome editing. J. Genet. Genomics 2017, 44, 175. [Google Scholar] [CrossRef]
  137. Qi, W.; Zhu, T.; Tian, Z.; Li, C.; Zhang, W.; Song, R. High-efficiency CRISPR/Cas9 multiplex gene editing using the glycine tRNA-processing system-based strategy in maize. BMC Biotechnol. 2016, 16, 58. [Google Scholar] [CrossRef] [Green Version]
  138. Andersson, M.; Turesson, H.; Nicolia, A.; Fält, A.S.; Samuelsson, M.; Hofvander, P. Efficient targeted multiallelic mutagenesis in tetraploid potato (Solanum tuberosum) by transient CRISPR-Cas9 expression in protoplasts. Plant Cell Rep. 2017, 36, 117–128. [Google Scholar] [CrossRef] [Green Version]
  139. Li, A.; Jia, S.; Yobi, A.; Ge, Z.; Sato, S.J.; Zhang, C.; Angelovici, R.; Clemente, T.E.; Holding, D.R. Editing of an alpha-kafirin gene family increases digestibility and protein quality in sorghum. Plant Physiol. 2018, 177, 1425–1438. [Google Scholar] [CrossRef] [Green Version]
  140. Zhang, B.; Yang, X.; Yang, C.; Li, M.; Guo, Y. Exploiting the CRISPR/Cas9 System for Targeted Genome Mutagenesis in Petunia. Sci. Rep. 2016, 6, 20315. [Google Scholar] [CrossRef]
  141. Xu, Z.S.; Feng, K.; Xiong, A.S. CRISPR/Cas9-Mediated Multiply Targeted Mutagenesis in Orange and Purple Carrot Plants. Mol. Biotechnol. 2019, 61, 191–199. [Google Scholar] [CrossRef]
  142. Nakajima, I.; Ban, Y.; Azuma, A.; Onoue, N.; Moriguchi, T.; Yamamoto, T.; Toki, S.; Endo, M. CRISPR/Cas9-mediated targeted mutagenesis in grape. PLoS ONE 2017, 12, e0177966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kaur, N.; Alok, A.; Shivani; Kaur, N.; Pandey, P.; Awasthi, P.; Tiwari, S. CRISPR/Cas9-mediated efficient editing in phytoene desaturase (PDS) demonstrates precise manipulation in banana cv. Rasthali genome. Funct. Integr. Genomics 2018, 18, 89–99. [Google Scholar] [CrossRef] [PubMed]
  144. Tian, S.; Jiang, L.; Gao, Q.; Zhang, J.; Zong, M.; Zhang, H.; Ren, Y.; Guo, S.; Gong, G.; Liu, F.; et al. Efficient CRISPR/Cas9-based gene knockout in watermelon. Plant Cell Rep. 2017, 36, 399–406. [Google Scholar] [CrossRef] [PubMed]
  145. Charrier, A.; Vergne, E.; Dousset, N.; Richer, A.; Petiteau, A.; Chevreau, E. Efficient targeted mutagenesis in apple and first time edition of pear using the CRISPR-Cas9 system. Front. Plant Sci. 2019, 10, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Hsu, C.-T.; Cheng, Y.-J.; Yuan, Y.-H.; Hung, W.-F.; Cheng, Q.-W.; Wu, F.-H.; Lee, L.-Y.; Gelvin, S.B.; Lin, C.-S. Application of Cas12a and nCas9-activation-induced cytidine deaminase for genome editing and as a non-sexual strategy to generate homozygous/multiplex edited plants in the allotetraploid genome of tobacco. Plant Mol. Biol. 2019, 101, 355–371. [Google Scholar] [CrossRef]
  147. Yin, X.; Anand, A.; Quick, P.; Bandyopadhyay, A. Editing a Stomatal Developmental Gene in Rice with CRISPR/Cpf1. In Plant Genome Editing with CRISPR Systems; Humana Press: New York, NY, USA, 2019; pp. 257–268. [Google Scholar]
  148. Meyer, J.D.F.; Deleu, W.; Garcia-Mas, J.; Havey, M.J. Construction of a fosmid library of cucumber (Cucumis sativus) and comparative analyses of the eIF4E and eIF(iso)4E regions from cucumber and melon (Cucumis melo). Mol. Genet. Genomics 2008, 279, 473–480. [Google Scholar] [CrossRef]
  149. 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]
  150. Wu, J.; Yang, R.; Yang, Z.; Yao, S.; Zhao, S.; Wang, Y.; Li, P.; Song, X.; Jin, L.; Zhou, T.; et al. ROS accumulation and antiviral defence control by microRNA528 in rice. Nat. Plants 2017, 3, 16203. [Google Scholar] [CrossRef]
  151. Sinharoy, S.; Liu, C.; Breakspear, A.; Guan, D.; Shailes, S.; Nakashima, J.; Zhang, S.; Wen, J.; Torres-Jerez, I.; Oldroyd, G.; et al. A medicago truncatula cystathionine-β-synthase-like domain-containing protein is required for rhizobial infection and symbiotic nitrogen fixation. Plant Physiol. 2016, 170, 2204–2217. [Google Scholar] [CrossRef] [Green Version]
  152. 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] [PubMed] [Green Version]
  153. 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]
  154. Wang, L.; Chen, L.; Li, R.; Zhao, R.; Yang, M.; Sheng, J.; Shen, L. Reduced drought tolerance by CRISPR/Cas9-mediated SlMAPK3 mutagenesis in tomato plants. J. Agric. Food Chem. 2017, 65, 8674–8682. [Google Scholar] [CrossRef] [PubMed]
  155. Wang, X.; Tu, M.; Wang, D.; Liu, J.; Li, Y.; Li, Z.; Wang, Y.; Wang, X. CRISPR/Cas9-mediated efficient targeted mutagenesis in grape in the first generation. Plant Biotechnol. J. 2018, 16, 844–855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Vu, T.V.; Sivankalyani, V.; Kim, E.J.; Doan, D.T.H.; Tran, M.T.; Kim, J.; Sung, Y.W.; Park, M.; Kang, Y.J.; Kim, J.Y. Highly efficient homology-directed repair using CRISPR/Cpf1-geminiviral replicon in tomato. Plant Biotechnol. J. 2020. [Google Scholar] [CrossRef] [Green Version]
  157. Peng, A.; Chen, S.; Lei, T.; Xu, L.; He, Y.; Wu, L.; Yao, L.; Zou, X. Engineering canker-resistant plants through CRISPR/Cas9-targeted editing of the susceptibility gene CsLOB1 promoter in citrus. Plant Biotechnol. J. 2017, 15, 1509–1519. [Google Scholar] [CrossRef] [Green Version]
  158. 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]
  159. Erpen-Dalla Corte, L.; Mahmoud, L.M.; Moraes, T.S.; Mou, Z.; Grosser, J.W.; Dutt, M. Development of Improved Fruit, Vegetable, and Ornamental Crops Using the CRISPR/Cas9 Genome Editing Technique. Plants 2019, 8, 601. [Google Scholar] [CrossRef] [Green Version]
  160. 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]
  161. Zhang, M.; Cao, Y.; Wang, Z.; Wang, Z.Q.; Shi, J.; Liang, X.; Song, W.; Chen, Q.; Lai, J.; Jiang, C. A retrotransposon in an HKT1 family sodium transporter causes variation of leaf Na+ exclusion and salt tolerance in maize. New Phytol. 2018, 217, 1161–1176. [Google Scholar] [CrossRef] [Green Version]
  162. Korotkova, A.M.; Gerasimova, S.V.; Khlestkina, E.K. Current achievements in modifying crop genes using CRISPR/Cas system. Vavilovskii Zhurnal Genet. Selektsii 2019, 23, 29–37. [Google Scholar] [CrossRef]
  163. Nemudryi, A.A.; Valetdinova, K.R.; Medvedev, S.P.; Zakian, S.M. TALEN and CRISPR/Cas genome editing systems: Tools of discovery. Acta Naturae 2014, 6, 22. [Google Scholar] [CrossRef]
  164. Li, T.; Liu, B.; Chen, C.Y.; Yang, B. TALEN-Mediated Homologous Recombination Produces Site-Directed DNA Base Change and Herbicide-Resistant Rice. J. Genet. Genomics 2016, 43, 297–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ji, X.; Zhang, H.; Zhang, Y.; Wang, Y.; Gao, C. Establishing a CRISPR–Cas-like immune system conferring DNA virus resistance in plants. Nat. Plants 2015, 1, 15144. [Google Scholar] [CrossRef] [PubMed]
  166. Oliva, R.; Ji, C.; Atienza-Grande, G.; Huguet-Tapia, J.C.; Perez-Quintero, A.; Li, T.; Eom, J.S.; Li, C.; Nguyen, H.; Liu, B.; et al. Broad-spectrum resistance to bacterial blight in rice using genome editing. Nat. Biotechnol. 2019, 37, 1344–1350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. De Toledo Thomazella, D.P.; Brail, Q.; Dahlbeck, D.; Staskawicz, B.J. CRISPR-Cas9 mediated mutagenesis of a DMR6 ortholog in tomato confers broad-spectrum disease resistance. BioRxiv 2016, 1, 064824. [Google Scholar]
  168. Giraud, T.; Gladieux, P.; Gavrilets, S. Linking the emergence of fungal plant diseases with ecological speciation. Trends Ecol. Evol. 2010, 25, 387–395. [Google Scholar] [CrossRef] [Green Version]
  169. Ali, Z.; Ali, S.; Tashkandi, M.; Zaidi, S.S.E.A.; Mahfouz, M.M. CRISPR/Cas9-Mediated Immunity to Geminiviruses: Differential Interference and Evasion. Sci. Rep. 2016, 6, 26912. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, F.; Wang, C.; Liu, P.; Lei, C.; Hao, W.; Gao, Y.; Liu, Y.G.; Zhao, K. Enhanced rice blast resistance by CRISPR/ Cas9-Targeted mutagenesis of the ERF transcription factor gene OsERF922. PLoS ONE 2016, 11, e0154027. [Google Scholar] [CrossRef]
  171. Pyott, D.E.; Sheehan, E.; Molnar, A. Engineering of CRISPR/Cas9-mediated potyvirus resistance in transgene-free Arabidopsis plants. Mol. Plant Pathol. 2016, 17, 1276–1288. [Google Scholar] [CrossRef] [Green Version]
  172. Chandrasekaran, J.; Brumin, M.; Wolf, D.; Leibman, D.; Klap, C.; Pearlsman, M.; Sherman, A.; Arazi, T.; Gal-On, A. Development of broad virus resistance in non-transgenic cucumber using CRISPR/Cas9 technology. Mol. Plant Pathol. 2016, 17, 1140–1153. [Google Scholar] [CrossRef] [Green Version]
  173. Zhang, T.; Zheng, Q.; Yi, X.; An, H.; Zhao, Y.; Ma, S.; Zhou, G. Establishing RNA virus resistance in plants by harnessing CRISPR immune system. Plant Biotechnol. J. 2018, 16, 1415–1423. [Google Scholar] [CrossRef] [Green Version]
  174. Baltes, N.J.; Hummel, A.W.; Konecna, E.; Cegan, R.; Bruns, A.N.; Bisaro, D.M.; Voytas, D.F. Conferring resistance to geminiviruses with the CRISPR-Cas prokaryotic immune system. Nat. Plants 2015, 1, 15145. [Google Scholar] [CrossRef]
  175. Ali, Z.; Abulfaraj, A.; Idris, A.; Ali, S.; Tashkandi, M.; Mahfouz, M.M. CRISPR/Cas9-mediated viral interference in plants. Genome Biol. 2015, 16, 238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Nekrasov, V.; Wang, C.; Win, J.; Lanz, C.; Weigel, D.; Kamoun, S. Rapid generation of a transgene-free powdery mildew resistant tomato by genome deletion. Sci. Rep. 2017, 7, 482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. 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]
  178. Iqbal, Z.; Sattar, M.N.; Shafiq, M. CRISPR/Cas9: A tool to circumscribe cotton leaf curl disease. Front. Plant Sci. 2016, 7, 475. [Google Scholar] [CrossRef] [Green Version]
  179. Sun, Q.; Lin, L.; Liu, D.; Wu, D.; Fang, Y.; Wu, J.; Wang, Y. CRISPR/Cas9-mediated multiplex genome editing of the BnWRKY11 and BnWRKY70 genes in brassica napus L. Int. J. Mol. Sci. 2018, 19, 2716. [Google Scholar] [CrossRef] [Green Version]
  180. Xu, Y.; Wang, F.; Chen, Z.; Wang, J.; Li, W.Q.; Fan, F.; Tao, Y.; Zhao, L.; Zhong, W.; Zhu, Q.H.; et al. Intron-targeted gene insertion in rice using CRISPR/Cas9: A case study of the Pi-ta gene. Crop J. 2019, 8, 424–431. [Google Scholar] [CrossRef]
  181. Zhang, Y.; Bai, Y.; Wu, G.; Zou, S.; Chen, Y.; Gao, C.; Tang, D. Simultaneous modification of three homoeologs of TaEDR1 by genome editing enhances powdery mildew resistance in wheat. Plant J. 2017, 91, 714–724. [Google Scholar] [CrossRef] [Green Version]
  182. Ortigosa, A.; Gimenez-Ibanez, S.; Leonhardt, N.; Solano, R. Design of a bacterial speck resistant tomato by CRISPR/Cas9-mediated editing of SlJAZ2. Plant Biotechnol. J. 2019, 17, 665–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Zhang, Z.; Ge, X.; Luo, X.; Wang, P.; Fan, Q.; Hu, G.; Xiao, J.; Li, F.; Wu, J. Simultaneous editing of two copies of GH14-3-3D confers enhanced transgene-clean plant defense against Verticillium dahliae in allotetraploid upland cotton. Front. Plant Sci. 2018, 9, 842. [Google Scholar] [CrossRef] [PubMed]
  184. Watson, A.; Ghosh, S.; Williams, M.J.; Cuddy, W.S.; Simmonds, J.; Rey, M.D.; Asyraf Md Hatta, M.; Hinchliffe, A.; Steed, A.; Reynolds, D.; et al. Speed breeding is a powerful tool to accelerate crop research and breeding. Nat. Plants 2018, 4, 23–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Tanksley, S.D.; Young, N.D.; Paterson, A.H.; Bonierbale, M.W. RFLP mapping in piant breeding: New tools for an old science. Bio/Technology 1989, 7, 257–264. [Google Scholar]
  186. Varshney, R.K.; Tuberosa, R. Genomics-Assisted Crop Improvement; Springer: Berlin/Heidelberg, Germany, 2007; Volume 2. [Google Scholar]
  187. Godwin, I.D.; Rutkoski, J.; Varshney, R.K.; Hickey, L.T. Technological perspectives for plant breeding. Theor. Appl. Genet. 2019, 132, 555–557. [Google Scholar] [CrossRef] [Green Version]
  188. Chiurugwi, T.; Kemp, S.; Powell, W.; Hickey, L.T.; Powell, W. Speed breeding orphan crops. Theor. Appl. Genet. 2018, 132, 607–616. [Google Scholar] [CrossRef]
  189. Cobb, J.N.; Biswas, P.S.; Platten, J.D. Back to the future: Revisiting MAS as a tool for modern plant breeding. Theor. Appl. Genet. 2019, 132, 647–667. [Google Scholar] [CrossRef] [Green Version]
  190. Andorf, C.; Beavis, W.D.; Hufford, M.; Smith, S.; Suza, W.P.; Wang, K.; Woodhouse, M.; Yu, J.; Lübberstedt, T. Technological advances in maize breeding: Past, present and future. Theor. Appl. Genet. 2019, 132, 817–849. [Google Scholar] [CrossRef] [Green Version]
  191. Mace, E.; Innes, D.; Hunt, C.; Wang, X.; Tao, Y.; Baxter, J.; Hassall, M.; Hathorn, A.; Jordan, D. The Sorghum QTL Atlas: A powerful tool for trait dissection, comparative genomics and crop improvement. Theor. Appl. Genet. 2019, 132, 751–766. [Google Scholar] [CrossRef]
  192. Monat, C.; Schreiber, M.; Stein, N.; Mascher, M. Prospects of pan-genomics in barley. Theor. Appl. Genet. 2019, 132, 785–796. [Google Scholar] [CrossRef]
  193. Varshney, R.K.; Hoisington, D.A.; Tyagi, A.K. Advances in cereal genomics and applications in crop breeding. Trends Biotechnol. 2006, 24, 490–499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Varshney, R.K.; Pandey, M.K.; Bohra, A.; Singh, V.K.; Thudi, M.; Saxena, R.K. Toward the sequence-based breeding in legumes in the post-genome sequencing era. Theor. Appl. Genet. 2019, 132, 797–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Ghosh, S.; Watson, A.; Gonzalez-Navarro, O.E.; Ramirez-Gonzalez, R.H.; Yanes, L.; Mendoza-Suárez, M.; Simmonds, J.; Wells, R.; Rayner, T.; Green, P.; et al. Speed breeding in growth chambers and glasshouses for crop breeding and model plant research. Nat. Protoc. 2018, 13, 2944–2963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Samineni, S.; Sen, M.; Sajja, S.B.; Gaur, P.M. Rapid generation advance (RGA) in chickpea to produce up to seven generations per year and enable speed breeding. Crop J. 2019, 8, 164–169. [Google Scholar] [CrossRef]
  197. Hong, W.; Kim, Y.; Kim, E.; Kumar, A.N.C.; Moon, S.; Gho, Y.; Yoou, M.; Kim, S.T.; Jung, K. CAFRI-Rice: CRISPR Applicable Functional Redundancy Inspector to Accelerate Functional Genomics in Rice. Plant J. 2020. [Google Scholar] [CrossRef]
  198. Li, J.; Manghwar, H.; Sun, L.; Wang, P.; Wang, G.; Sheng, H.; Zhang, J.; Liu, H.; Qin, L.; Rui, H.; et al. Whole genome sequencing reveals rare off-target mutations and considerable inherent genetic or/and somaclonal variations in CRISPR/Cas9-edited cotton plants. Plant Biotechnol. J. 2019, 17, 858–868. [Google Scholar] [CrossRef]
  199. Hajiahmadi, Z.; Movahedi, A.; Wei, H.; Li, D.; Orooji, Y.; Ruan, H.; Zhuge, Q. Strategies to increase on-target and reduce off-target effects of the CRISPR/Cas9 system in plants. Int. J. Mol. Sci. 2019, 20, 3719. [Google Scholar] [CrossRef] [Green Version]
  200. Lu, Y.; Tian, Y.; Shen, R.; Yao, Q.; Wang, M.; Chen, M.; Dong, J.; Zhang, T.; Li, F.; Lei, M.; et al. Targeted, efficient sequence insertion and replacement in rice. Nat. Biotechnol. 2020. [Google Scholar] [CrossRef]
  201. Rodríguez-Leal, D.; Lemmon, Z.H.; Man, J.; Bartlett, M.E.; Lippman, Z.B. Engineering quantitative trait variation for crop improvement by genome editing. Cell 2017, 171, 470–480. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Overview of the different aspects covered in the present review related to genome editing in plants, such as its applications, challenges, and advantages.
Figure 1. Overview of the different aspects covered in the present review related to genome editing in plants, such as its applications, challenges, and advantages.
Ijms 21 05665 g001
Figure 2. Processes of Zinc-finger nucleases (ZFNs). (a) Delivery of ZFNs into cells by transfection or electroporation. (b) ZFNs are fusions of the nonspecific DNA cleavage domain of the FokI restriction nuclease with zinc-finger proteins. ZFN dimers induce targeted and double-stranded breaks (DSBs) to stimulate the DNA damage response pathways. The binding specificity of the designed zinc-finger domain directs the ZFN to a particular genomic site. The boxes with blue color represent the mechanism of the ZFN genome-editing technique. (c) Cloning of ZFN-genome-edited cells and screening of positive clones by RT-PCR and sequencing analyses. This figure briefly modified from the source (https://www.sigmaaldrich.com/technical-documents/articles/life-science-innovations/targeted-genome-editing.html).
Figure 2. Processes of Zinc-finger nucleases (ZFNs). (a) Delivery of ZFNs into cells by transfection or electroporation. (b) ZFNs are fusions of the nonspecific DNA cleavage domain of the FokI restriction nuclease with zinc-finger proteins. ZFN dimers induce targeted and double-stranded breaks (DSBs) to stimulate the DNA damage response pathways. The binding specificity of the designed zinc-finger domain directs the ZFN to a particular genomic site. The boxes with blue color represent the mechanism of the ZFN genome-editing technique. (c) Cloning of ZFN-genome-edited cells and screening of positive clones by RT-PCR and sequencing analyses. This figure briefly modified from the source (https://www.sigmaaldrich.com/technical-documents/articles/life-science-innovations/targeted-genome-editing.html).
Ijms 21 05665 g002
Figure 3. Transcription activator-like effector nucleases (TALENs) are dimeric transcription factors/nucleases engineered from an array of 34-amino-acid molecules, each of which targets one nucleotide. The target sequence is recognized; a corresponding TALEN sequence is built and inserted into a cellular plasmid. The cellular plasmid is inserted into the host cell, where it is translated to produce the functional TALEN, which penetrates the nucleus and binds to and cleaves the target sequence. The applications of this system include the knockout of a target gene or the addition of a replacement nucleotide into the target gene.
Figure 3. Transcription activator-like effector nucleases (TALENs) are dimeric transcription factors/nucleases engineered from an array of 34-amino-acid molecules, each of which targets one nucleotide. The target sequence is recognized; a corresponding TALEN sequence is built and inserted into a cellular plasmid. The cellular plasmid is inserted into the host cell, where it is translated to produce the functional TALEN, which penetrates the nucleus and binds to and cleaves the target sequence. The applications of this system include the knockout of a target gene or the addition of a replacement nucleotide into the target gene.
Ijms 21 05665 g003
Figure 4. CRISPR/Cas9-based genome editing. (a) Selection of the desired target on genomic DNA and recognition of protospacer adjacent motif (PAM) sequences before 20-base-pair sequences. (b) Design of the sgRNA using different online bioinformatics tools. (c) Subsequent step toward the cloning of designed sgRNAs and the construction of the binary vector using different promoters. (d) Transfer of the vector into the plant by Agrobacterium tumefaciens-mediated plant transformation and development of transgenic plants. (e) Genotyping analysis of transgenic plants, as mentioned in the figure.
Figure 4. CRISPR/Cas9-based genome editing. (a) Selection of the desired target on genomic DNA and recognition of protospacer adjacent motif (PAM) sequences before 20-base-pair sequences. (b) Design of the sgRNA using different online bioinformatics tools. (c) Subsequent step toward the cloning of designed sgRNAs and the construction of the binary vector using different promoters. (d) Transfer of the vector into the plant by Agrobacterium tumefaciens-mediated plant transformation and development of transgenic plants. (e) Genotyping analysis of transgenic plants, as mentioned in the figure.
Ijms 21 05665 g004
Figure 5. In a Cpf1-mediated plant genome-editing system, the T-rich region (TTTN) acts as a protospacer adjacent motif (PAM). (a,b) Cpf1 cleaves the target DNA and introduces double-stranded breaks (DSBs), a 5-nt potential staggered cut distal to a 5′ T-rich PAM. (c,d) In Cpf1, the DSBs are subsequently repaired by two primary cellular mechanisms, nonhomologous end joining (NHEJ) and homology-directed repair (HDR).
Figure 5. In a Cpf1-mediated plant genome-editing system, the T-rich region (TTTN) acts as a protospacer adjacent motif (PAM). (a,b) Cpf1 cleaves the target DNA and introduces double-stranded breaks (DSBs), a 5-nt potential staggered cut distal to a 5′ T-rich PAM. (c,d) In Cpf1, the DSBs are subsequently repaired by two primary cellular mechanisms, nonhomologous end joining (NHEJ) and homology-directed repair (HDR).
Ijms 21 05665 g005
Figure 6. Case studies using the speed editing strategy for gene-family members: Rice contains 33,483 Pfam annotated genes. Among them, the functional significance of homologous genes within a family can be evaluated by integrated transcriptome data and Pearson’s correlation coefficient. OsMADS62 and OsMAD63 in the same sisternode of the phylogenetic tree showed redundant expression in mature pollen with the highest expression, and their double mutant exhibited only a male sterile phenotype via a multiple CRISPR/Cas9 system. Conversely, RUPO showed predominant expression over LOC_Os03g55210 in mature pollen, and a single mutation of the RUPO gene via the CRISPR/Cas9 system caused a male sterile phenotype.
Figure 6. Case studies using the speed editing strategy for gene-family members: Rice contains 33,483 Pfam annotated genes. Among them, the functional significance of homologous genes within a family can be evaluated by integrated transcriptome data and Pearson’s correlation coefficient. OsMADS62 and OsMAD63 in the same sisternode of the phylogenetic tree showed redundant expression in mature pollen with the highest expression, and their double mutant exhibited only a male sterile phenotype via a multiple CRISPR/Cas9 system. Conversely, RUPO showed predominant expression over LOC_Os03g55210 in mature pollen, and a single mutation of the RUPO gene via the CRISPR/Cas9 system caused a male sterile phenotype.
Ijms 21 05665 g006
Figure 7. The illustration represents the major strategies aimed at improving the genome-editing systems.
Figure 7. The illustration represents the major strategies aimed at improving the genome-editing systems.
Ijms 21 05665 g007
Table 1. Comparison of EMNs, ZFNs, TALENs, CRISPR/Cas9, and CRIPSR/Cpf1.
Table 1. Comparison of EMNs, ZFNs, TALENs, CRISPR/Cas9, and CRIPSR/Cpf1.
FunctionsEMNsZFNsTALENsCRIPSRs/Cas9CRIPSRs/Cpf1References
Mode of actionInformation strand directs conversion(s) within the target regionDouble-strand breaks in the target DNADouble-strand breaks in the target DNADouble-strand breaks or single-strand nicks in the target DNADouble-strand breaks[56,57,58,59]
Target recognition
efficiency
HighHighHighHighVery High[59,60]
Mutation rateMiddleHighMiddleLowHigh[4,56,59]
Creation of large-scale librariesTechnically difficultImpossibleTechnically difficultPossiblePossible[59,61,62]
MultiplexingTechnically difficultDifficultDifficultPossiblePossible[56,57,59]
ComponentsExogenous polynucleotide
(chimeraplast)
Zn finger domains
Nonspecific FokI nuclease domain
TALE DNA-binding domains Nonspecific FokI nuclease domaincrRNA, Cas9
proteins
crRNA, Cpf1
proteins
[59,60,63]
Structural
protein
Dimeric proteinDimeric proteinDimeric proteinMonomeric
Protein
Monomeric
Protein
[4,56,59]
Catalytic
Domain
Absence of a catalytic domainRestriction endonuclease FokIRestriction endonuclease FokIRuvC and HNHRuvC and HNH[59,63,64]
Length of the target sequence (bp)68–8824–3624–5920–2220–24[4,61,65]
Protein
engineering
steps
Not requiredRequiredRequiredShould not be difficult to test gRNAShould not be difficult to test gRNA[59,62,66]
CloningNot necessaryNecessaryNecessaryNot necessaryNot necessary[59,62,66]
gRNA
production
Not requiredNot applicableNot applicableEasy to produceEasy to produce[59,62,67]
Target genome-editing toolsNot RequiredZFNGenome v2.0
ZifBASE
Zinc-Finger
Database
(ZiFDB)
Zinc-Finger Tool
EENdb
TALE-NT 2.0
SPATA
TALEN offer
TALEN Library
T
CHOP CHOP
CRISPRs web Server
Crass: The CRISPR Assembler CRISPR Target
Breaking-Cas
Cas-OFFinder
CRISPOR
CCTOP
[46,68]
Off-target
effects
Low off-target effectLow off-target effectShows least off-target activitiesLow off-target effectLow off-target effect[69]
Cost of developmentHighHighHigherLowLow[63,70,71]
Table 2. Improvement in crops, fruits, and vegetables via EMNs, ZFNs, and TALENs.
Table 2. Improvement in crops, fruits, and vegetables via EMNs, ZFNs, and TALENs.
ToolsCrop/Fruits/VegetableTarget GeneTrait ImprovementReferences
EMNsMaizeMS26Independent lines of male sterile plants[88]
EMNsCottonEPSPSHerbicide tolerance[89]
ZFNSoybeanDCLHerbicide transmission[77]
ZFNMaizePATHerbicide resistance[29]
ZFNTobaccoGUS: NPTIIChromosome breaks[73]
ZFNRICEOsQQRDetection of safe harbor loci
Herbicide
[90]
TALENWheatTaMLOPowdery mildew resistance[91]
TALENPotatoEndogenous consist. PromoterHerbicide resistance[4]
TALENPotatoALSHerbicide resistance[92]
TALENPotatoVacuolar invertaseNo reducing sugars and improved food safety[93]
TALENSugarcaneCaffeic acid O-methyltransferaseReduced lignin and improved biofuel production[94]
TALENPotatoVlnvLow concentration of reducing sugars and undetectable concentration of reducing sugars[93]
TALENRiceOsBADH2Fragrant rice[95]
TALENSoybeanFAD2-1A, FAD2-1BLow polyunsaturated fats[96]
TALENWheatTaMLO-A1, TaMLO-B1, TaMLO-D1Powdery mildew resistance[91]
Table 3. Improvement in the yield and quality of crops, fruits, and vegetables via CRISPR.
Table 3. Improvement in the yield and quality of crops, fruits, and vegetables via CRISPR.
ToolsCrop/Fruit/VegetableTarget GeneTrait ImprovementReferences
CRISPR/Cas9RiceGn1a, GS3, and DEP1Grain number, grain size, panicle architecture[129,130]
CRISPR/Cas9WheatTaGASR7Grain length and weight[131]
CRISPR/Cas9FlaxFAD2Seed oil composition (high oleic and low polyunsaturated FAs)[97]
CRISPR/Cas9SoybeanGmFT2aLate flowering[132]
CRISPR/Cas9TomatoSP5GTime to harvest[116]
CRISPR/Cas9TomatoRINFruit ripening (shelf life)[133]
CRISPR/Cas9TomatoSlIAA9Parthenocarpy (leading to seedless fruit)[115]
CRISPR/Cas9WheatPDSChlorophyll syn[57]
CRISPR/Cas9CottonALARPCotton fiber development[134]
CRISPR/Cas9RiceWaxyEnhanced glutinosity[135]
CRISPR/Cas9RiceHd2, Hd4, Hd5Early heading[136]
CRISPR/Cas9Maize PPR, RPLReduced zein protein[137]
CRISPR/Cas9PotatoGBSSIncreased amylopectin/amylose[138]
CRISPR/Cas9SorghumWholek1Cgene familyIncrease in the grain protein digestibility and lysine content[139]
CRISPR/Cas9PetuniaPDSThe biosynthesis of carotenoid and chlorophyll[140]
CRISPR/Cas9CarrotDcPDS, DcMYB113Purple depigmented carrot[141]
CRISPR/Cas 9CabbageBolc.GA4.aDwarfing and fruit dehiscence[108]
CRISPR/Cas9GrapeVvPDS, MLO-7Albino phenotype[142]
CRISPR/Cas 9BananaPDSAlbino and variegated phenotype[143]
CRISPR/Cas 9WatermelonClPDSAlbino phenotype[144]
CRISPR/Cas9ApplePDS, TFL1Albino phenotype, early flowering[145]
CRISPR/Cpf1TobaccoETR1Plants harboring[146]
CRISPR/Cpf1MaizePAP1Stable mRNA equal[47]
CRISPR/Cpf1RiceOsROC5, OsDEP1Mutation frequencies doubled[47]
CRISPR/Cpf1RiceOsEPFL9Regulation of stomatal density[147]
Table 4. Improvement of climate-resilient crops, vegetables, and fruits by CRISPR/Cas9/Cpf1.
Table 4. Improvement of climate-resilient crops, vegetables, and fruits by CRISPR/Cas9/Cpf1.
ToolsCrop/Fruit/VegetableTarget GeneTrait ImprovementReferences
CRISPR/Cas9MaizeARGOS8Drought tolerance[157]
CRISPR/Cas9RiceOsNAC041Salinity tolerance[158]
CRISPR/Cas9TomatoNPRIDrought tolerance[159]
CRISPR/Cas9SoybeanDrb2a and Drb2bSalt and drought tolerance[153]
CRISPR/Cas9TomatoSIMAPK3Drought tolerance[154]
CRISPR/Cas9TomatoSIAGL6Heat stress[123]
CRISPR/Cas9GrapesWRKY52,Biotic stress responses[155]
CRISPR/Cas9SoybeanSAPK1 and SAPK2Salinity tolerance[160]
CRISPR/Cas9MaizeZmHKT1Salinity tolerance[161]
CRISPR/Cas9RiceOsMPK2, OsPDS, OsBADH2Multiple-stress tolerance[162]
CRISPR/Cpf1TomatoHKT1;2 HDRMultiple-stress tolerance[156]
Table 5. Improvement of plant disease resistance by CRISPR in crops, fruits, and vegetables.
Table 5. Improvement of plant disease resistance by CRISPR in crops, fruits, and vegetables.
ToolsCrop/Fruit/VegetableTarget GeneTrait ImprovementReferences
CRISPR/Cas 9Citrus (orange)CsLOB1 (promoter)Citrus canker resistance[157]
CRISPR/Cas 9CucumbereIF4EBroad virus resistance[172]
CRISPR/Cas 9Tobacco43 regions in the viral genomeResistance to the Gemini virus beet severe curly top virus[173]
CRISPR/Cas 9TobaccoSix regions in the viral genomeResistance to the Gemini virus bean yellow dwarf virus[174]
CRISPR/Cas 9TomatoThree regions in the viral genomeResistance to the Gemini virus Resistance to the tomato yellow leaf curl virus[175]
CRISPR/Cas 9TomatoSlMlo1Resistance to powdery mildew[176]
CRISPR/Cas 9WheatMLO-A1, TaMLO-B1 and TaMLO-D1Resistance to powdery mildew[91]
CRISPR/Cas9GrapeVvPDS, MLO-7Albino phenotype
Powdery mildew resistance
[120,142]
CRISPR/Cas9WheatTaMLOPowdery mildew resistance[91]
CRISPR/Cas9PotatoS-genesPhytophthora infestans resistance[138]
CRISPR/Cas9CottonViral and satellite DNAsResistance to cotton leaf curl disease[177]
CRISPR/Cas9CitrusCsLOB1Canker resistance[178]
CRISPR/Cas9AppleDIPM-1, DIPM-2, and DIPM-4 genesResistance to fire blight disease[120]
CRISPR/Cas9PotatoS-genesPhytophthora infestans resistance[138]
CRISPR/Cas9RapeseedWRKY70, WRKY11JA- and SA-induced resistance to pathogens[179]
CRISPR/Cas9RicePi-taResistance to the rice blast disease[180]
CRISPR/Cas9WheatEDR1Improved resistance
against powdery mildew
[181]
CRISPR/Cas9TomatoSlJAZ2Bacterial speck resistance[182]
CRISPR/Cas9CottonGh14-3-3Resistance to cotton verticillium wilt[183]

Share and Cite

MDPI and ACS Style

Ahmar, S.; Saeed, S.; Khan, M.H.U.; Ullah Khan, S.; Mora-Poblete, F.; Kamran, M.; Faheem, A.; Maqsood, A.; Rauf, M.; Saleem, S.; et al. A Revolution toward Gene-Editing Technology and Its Application to Crop Improvement. Int. J. Mol. Sci. 2020, 21, 5665. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21165665

AMA Style

Ahmar S, Saeed S, Khan MHU, Ullah Khan S, Mora-Poblete F, Kamran M, Faheem A, Maqsood A, Rauf M, Saleem S, et al. A Revolution toward Gene-Editing Technology and Its Application to Crop Improvement. International Journal of Molecular Sciences. 2020; 21(16):5665. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21165665

Chicago/Turabian Style

Ahmar, Sunny, Sumbul Saeed, Muhammad Hafeez Ullah Khan, Shahid Ullah Khan, Freddy Mora-Poblete, Muhammad Kamran, Aroosha Faheem, Ambreen Maqsood, Muhammad Rauf, Saba Saleem, and et al. 2020. "A Revolution toward Gene-Editing Technology and Its Application to Crop Improvement" International Journal of Molecular Sciences 21, no. 16: 5665. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21165665

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