Next Article in Journal
Cell Growth Inhibition, DNA Fragmentation and Apoptosis-Inducing Properties of Household-Processed Leaves and Seeds of Fenugreek (Trigonella Foenum-Graecum Linn.) against HepG2, HCT-116, and MCF-7 Cancerous Cell Lines
Next Article in Special Issue
Analysis of the Chloroplast Genome of Ficus simplicissima Lour Collected in Vietnam and Proposed Barcodes for Identifying Ficus Plants
Previous Article in Journal
Evaluation of Antidiabetic, Antioxidant and Anti-Hyperlipidemic Effects of Solanum indicum Fruit Extract in Streptozotocin-Induced Diabetic Rats
Previous Article in Special Issue
BSR and Full-Length Transcriptome Approaches Identified Candidate Genes for High Seed Ratio in Camellia vietnamensis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications and Prospects of CRISPR/Cas9-Mediated Base Editing in Plant Breeding

1
College of Agronomy, Anhui Agricultural University, Heifei 230036, China
2
Key Laboratory of Rice Genetic Breeding of Anhui Province, Rice Research Institute, Anhui Academy of Agricultural Sciences, Hefei 230031, China
*
Authors to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(2), 918-935; https://0-doi-org.brum.beds.ac.uk/10.3390/cimb45020059
Submission received: 24 December 2022 / Revised: 10 January 2023 / Accepted: 17 January 2023 / Published: 19 January 2023
(This article belongs to the Special Issue Functional Genomics and Comparative Genomics Analysis in Plants)

Abstract

:
The clustered regularly interspaced short palindromic repeats (CRISPR)/associated protein 9 system (Cas9) has been used at length to optimize multiple aspects of germplasm resources. However, large-scale genomic research has indicated that novel variations in crop plants are attributed to single-nucleotide polymorphisms (SNPs). Therefore, substituting single bases into a plant genome may produce desirable traits. Gene editing by CRISPR/Cas9 techniques frequently results in insertions–deletions (indels). Base editing allows precise single-nucleotide changes in the genome in the absence of double-strand breaks (DSBs) and donor repair templates (DRTs). Therefore, BEs have provided a new way of thinking about genome editing, and base editing techniques are currently being utilized to edit the genomes of many different organisms. As traditional breeding techniques and modern molecular breeding technologies complement each other, various genome editing technologies have emerged. How to realize the greater potential of BE applications is the question we need to consider. Here, we explain various base editings such as CBEs, ABEs, and CGBEs. In addition, the latest applications of base editing technologies in agriculture are summarized, including crop yield, quality, disease, and herbicide resistance. Finally, the challenges and future prospects of base editing technologies are presented. The aim is to provide a comprehensive overview of the application of BE in crop breeding to further improve BE and make the most of its value.

1. Introduction

The clustered regularly interspaced, short palindromic repeats (CRISPR)/CRISPR-associated 9 (Cas9) is a third-generation gene editing technology following ZFNs and TALENs. It has the advantages of being highly efficient, simple, inexpensive, and easily usable [1,2,3].
In the CRISPR/Cas9 system, a Cas9-single guide RNA (sgRNA) complex binds to a specific nucleotide sequence with the guidance of the sgRNA and cleaves the target DNA strand, causing a double-strand break (DSB) [4,5,6,7]. These DSBs can be corrected by nonhomologous end-joining (NHEJ) or the homology-directed repair (HDR) mechanism [8,9]. NHEJ is a method of repair in which the ends of DSBs are directly linked by DNA ligase and do not depend on homologous DNA sequences; therefore, NHEJ repair is rapid but not exact. The homologous repair process is complex and precise but requires a homologous DNA sequence template and can occur only in the G2/S phase of a cell [10,11,12,13].
Due to the genetic basis underlying the diversity of many important crop species and single-nucleotide variations [14,15], it is necessary to develop a technique that allows precise and effective single-base substitutions. Base editing technology is a novel target gene modification technique developed based on the CRISPR/Cas system, by its utilization of a tethered deaminase domain or nickase Cas9 for base conversion from A > G or C > T or C > G without the donor DNA and a DSB introduction in the genome. Recent studies have utilized Base-editors to create single and multiple nucleotide modifications in cells. Here we review different base editing technologies and their applications in crop improvement.

2. Base Editing

Base editors (BEs) enable single-nucleotide targeted mutations without severing the nucleic acid backbone and enable direct chemical modification of target nucleobases. The original base editor used a single-stranded DNA-specific cytidine deaminase combined with an inactivated Cas9 (dCas9) to convert a cytosine (C)-guanine (G) base pair to thymine (T)-adenine (A) in the target region, called the cytosine base editor (CBE). Later researchers developed the Adenine Base Editor and the Guanine Base Editor based on the CBE. Currently available base editing systems include cytosine (C) base editors (CBEs), adenine (A) base editors (ABEs), and guanine (G) base editors (CGBEs) (Figure 1) [16,17,18,19]. Each of these categories is discussed below.

3. CBEs

The first-generation CBE, BE1, consists of rat C deaminase (rAPOBEC1) and dCas9, whose cleavage activity is completely lost [20,21]. When the fusion protein targets genomic DNA under the guidance of sgRNA, C deaminase can bind to the ssDNA in the R-loop region formed by the Cas9 protein, sgRNA, and genomic DNA and deaminate C to uracil (U) within a certain range along the ssDNA. During DNA replication, U is read by DNA polymerase as thymine (T). The final substitution of C/G to T/A base pairs then occurs (Figure 1a). Later researchers developed a second-generation cytosine base editor, BE2, by incorporating a uracil DNA glycosylase inhibitor (UGI) from phage PBS on top of BE1. Because UGI can inhibit the action of uracil DNA glycosylase (UDG) in the organism, BE2 is three times more efficient at editing than BE1. CBEs have undergone several generations of updates; notably, BE3 has replaced dCas9 in BE2 with nCas9(D10A) [22,23,24,25]. nCas9(D10A) specifically creates a gap in the nonedited strand, which in turn stimulates the intracellular base mismatch repair pathway (MMR) [26,27], which uses the editing strand containing U as a template for repair, resulting in increased editing efficiency. These optimized CBEs can better serve precision breeding [18,24,28]. However, existing CBEs/CGBEs rely on the natural cytosine deaminase AID/APOBEC and often produce high insertional deletion by-products and off-target effects due to the activation of the base excision repair pathway by cytosine deamination. Recently, researchers have transformed the adenine deaminase TadA-8e into a non-natural cytosine deaminase using only cytosine as a substrate and constructed the first novel CGBE/CBE family of base editors—Td-CGBE/Td-CBEs—that do not rely on the AID/APOBEC deaminase family, demonstrating lower off-target effects and very low indels events. Constructing a corresponding base editor in plants will facilitate crop improvement [29].

4. ABEs

Three main components compose ABEs: synthetic A deaminase, nCas9 (D10A), and sgRNA [30]. The A deaminase protein binds to ssDNA and deaminates A into inosine I, which is then read and replicated as G at the DNA level. This enables the instant exchange of A–T base pairs with G–C base pairs when the fusion protein targets genomic DNA under the guidance of sgRNA [31] (Figure 1b). Using ABEs eliminates the limitation that CBEs can edit only C or G and opens up a wider range of base transformation possibilities. In contrast to CBEs, ABEs do not require the suppression of alkyl adenine DNA glycosylase (AAG) activity [32,33,34].

5. CGBEs

C-to-G base editors (CGBEs) were constructed by adapting existing CBE tools to generate a new tool suitable for mediating C–G base reversals [35].
Combining a Cas9 nickase (nCas9-D10A), cytidine deaminase, and Uracil-N-glycosylase (UNG) leads to the production of CGBEs. Cytidine deaminase causes the conversion of a target C to U under the guidance of RNA. UNG locates U in the DNA and eliminates it, resulting in the formation of an AP site [36]. When nCas9 creates the AP site and binds the nonedited strand, DNA repair and replication mechanisms are triggered, preferentially inserting a G at the AP site. In contrast to CBEs, which contain a UNG inhibitor, CGBEs contain UNG [37,38] (Figure 1c).
In the base editor, the single-stranded DNA in the R-loop is exposed during base editing. This single-stranded DNA binds to the 20 bp of the sgRNA, but there is a preference for the action of cytidine deaminase on this 20 bp fragment so that different base editors have specific BE activity windows. We define the 20th base from the PAM site as the first position. The Activity Window of the Classical Cytosine Base Editor BE3-SpCas9 is bases 4–10. Bases 4–12 of the classical adenine is the base editor ABE7.10-SpCas9 activity window. Bases 5–7 of the classical guanine base editor is the CGBE-SpCas9 activity window. The different base editor activity windows depend on various factors, such as Cas proteins, deaminases, and variant connectors [39,40,41] (Figure 1d).

6. Application of Base Editors in Plants

In the long history of breeding, several major strategies have been used, such as crossbreeding, mutation breeding, etc. Gene editing and transgenics are an important part of the new breeding era [42,43,44,45,46,47]. Crossbreeding can only introduce known good traits [48,49]. Mutation breeding is a longer breeding process in which researchers create random mutations throughout the plant genome through physical and chemical mutagenesis. Transgenic breeding techniques allow for the direct introduction of good genes specific to a crop or genes from other species to obtain crop varieties with higher yields and better nutritional quality [50,51,52,53]. However, this breeding method requires the integration of exogenous genes into the plant genome and is, therefore, subject to strict controls. BE technology is an effective complement to the above three breeding methods. BEs allow for targeted modification of the plant genome without introducing exogenous genes to obtain the target variety quickly [54,55,56,57].
In recent years, public investment in research has been used to sequence, assemble and annotate the genomes of major crops, and we have gained a wealth of functional genetic information on plants [54,55,56,57,58]. Base editing allows for precise genome editing, and its successful operation in breeding has opened up new opportunities for crop improvement. Since 2016, BEs have been used to edit the genomes of various plant species, including rice, maize, cotton, oilseed rape, tomato, strawberry, and watermelon [59,60,61,62,63,64,65,66,67,68,69]. The contributions of BEs to improved yields, increased stress resistance, improved herbicide resistance, and quality-regulated nutrient composition of various crop species are reviewed in this paper. As traditional breeding techniques and modern molecular breeding technologies complement each other, various genome editing technologies have emerged [70,71,72]. The question of how to realize the greater potential of BE applications is one that we need to consider.

7. Increasing Yields

Satisfying the growing demand for increased crop yields is extremely challenging due to soil degradation, climate change, and many other constraints [73,74]. In recent years, traditional breeding has been dedicated to increasing crop yields. However, traditional breeding is accompanied by long cycle times and often by linkage resistance. Many plant genetic variants contain deleterious mutations. Mutation breeding, on the other hand, requires multigenerational crosses, backcrosses, etc. The process is time-consuming and laborious. BEs allow for precisely targeted modifications to plant genomes and can rapidly alter species traits. OsSPL regulates the meiotic fate in rice. With this in mind, it is possible to generate high-yielding rice lines by developing various sgRNAs and different adenylation base codons that target OsSPL14, OsSPL16, OsSPL17, and OsSPL18 sequences [75,76] (Table S1). A point mutation in the OsmiR156 binding site of OsSPL14 leads to OsmiR156-mediated cleavage of the OsSPL14 transcript, resulting in rice plants with desirable architecture and increased seed yields [77,78,79] (Table S1). It was shown that base editing improved yields without introducing a cumbersome trait. The OsSPL14 (which governs the desired rice structure) and OsSIR16 (which controls rice grain size, shape, and quality) genes were simultaneously edited by the use of CBE(A3A/Y130F-CBE-V01) to generate lines with improved high expression of OsSPL14 and OsSPL16 (Table S1). Concerning these edited lines, we can obtain increased yields [75,80,81]. Base editing can also be used to produce higher-yielding crops using other strategies. CBE can produce C-to-T base mutations at specific sites, converting non-terminating codons to stop codons and thereby silencing gene expression. OsGS3 is a quantitative trait locus (QTL) that regulates the rice grain length and width; OsGW2 is a QTL that controls rice grain width and weight; regulation of the expression of OsGS3 and OsGW2 genes using CBE-driven insertion of premature stop codons can generate high-yielding rice lines [80,82,83,84] (Table S1). It follows that, in the case of known target genes, the application of BE in plant crops is a novel strategy to improve crop traits for more efficient and sustainable agriculture.

8. Improving Quality

The quality of crop products is directly related to the edibility and economic value of a crop. The amylose content (AC) is a critical determinant of the edible and culinary quality (ECQ) of rice [85,86]. Specifically, Xian (indica) and Geng (japonica), both of which are commonly consumed, comprise 10% to 20% straight-chain starch [87,88].
Rice with a medium-to-low AC concentration is often softer, with higher palatability and increased glossiness. Rice with a high AC is stiffer [89,90]. The rice Waxy (Wx) gene is vital for the edibility and culinary quality (ECQ) of rice. Wx encodes grain-bound starch synthase I (GBSSI), an enzyme that regulates the synthesis of straight-chain starch and ultimately determines the endosperm’s straight-chain starch content [91,92,93]. The GBSS allele of Xian rice variety YK17 was edited via ABEs to decrease the AC content while maintaining ECQ parameters such as gel consistency (GC) and alkali diffusion value; these traits could significantly enhance the ECQ of Xian rice variety YK17 [94] (Table S1). Similar methods were used to produce mutant Geng rice lines with a low AC content, where precise BE of the Geng Wx gene by CBEs controlled the overall abundance and activity of GBSSI. Mutant Geng rice with a low AC content was obtained [92] (Table S1). In addition, three starch branching enzymes (SBE), OsSBEI, OsSBEIIa, and OsSBEIIb, are included in rice. OsSBEIIb is specifically expressed in the endosperm and can transfer short chains to branched starch crystals. Inactivation of OsSBEIIb leads to reduced branching of branched starch, increased straight-chain starch content, increased seed opacity, and reduced dry weight [95,96,97]. Researchers selected two targets (S3 and S5) within the rice OsSBEIIb gene and successfully induced point mutations via pCXUN-BE3 (Table S1); this experiment highlighted a viable and effective tool for the modification of straight-chain starch content in rice.
BEs can be applied to other crop species and also improve traits in economically important crop species. Tomato fruit and tomato-related foods provide a significant amount of lycopene to humans [98,99,100]. The primary pigments in tomato fruit are carotenoids, such as lycopene and beta-carotene. Carotenoids are essential functional components because of their strong antioxidant ability, and SlDDB1, SlDET1, and SlCYC-B play significant roles in carotenoid accumulation. Target-AID, developed in 2016, can introduce a base substitution within a target gene. Target-AID binds PmCDA1 and dCas9 from the eel. The basic principle of Target-AID is the same as BE, but the deaminase used is different; the deaminase used in Target-AID is PmCDA1, which is in the AID family [66,101,102,103]. Targeting studies on the SlDDB1, SlDET1, and SlCYC-B genes have been conducted using the Target-AID technique. The findings suggested that the BE strategy could increase the carotenoid content of tomato fruit and that the altered lines presented significant changes in carotenoid accumulation [66] (Table S1). In addition, strawberry is another of the major economically important crop species. Different genotypes of strawberries had varying effects on the sugar content of their fruit, thus enhancing the genotype and trait variety. Human APOBEC3A, coupled with the Cas9 protein to increase plant base editing effectiveness, was used to generate the novel plant BE known as A3A-PBE [104]. Researchers have utilized A3A-PBE to edit the conserved uORF of the strawberry FvebZIPs1.1 gene to generate seven novel uORFs in the T0 generation. The homozygous mutants with the seven novel uORF mutations presented varying degrees of increased fruit sugar contents, and there was no effect on plant growth (Table S1). This example shows that, using the Single Base Editor, it is possible to precisely adjust the sugar content of strawberries for different production requirements [67].

9. Crop Morphology and Nitrogen Uptake

In addition, BE can not only be used to improve the nutrient content of crop plants but can also enhance crop morphology and regulate nitrogen uptake. NRT1.1B encodes a nitrogen transporter [105,106] (Table S1), and research has indicated that a C→T substitution (Thr327Met) in this gene increases nitrogen use efficiency in rice. SLR1 encodes a DELLA protein with an amino acid substitution in or near its TVHYNP sequence, reducing plant height [107,108,109] (Table S1). Similarly, CBEs have been used to target specific regions of IAA7 and RGA, which encode growth hormone response proteins and gibberellin signaling proteins, resulting in targeted mutations of C bases to T bases at the target site. Plants exhibited a dwarf-type morphology to varying degrees after mutations of single bases in IAA7 and RGA (Table S1). In addition, another study showed that base editing systems could be used for the genetic improvement of kale-type oilseed rape adapted to mechanized breeding [110] (Table S1). It follows that base editing can be used for crop improvement and germplasm innovation. We can use base editing to construct sgRNA libraries to screen for mutant loci that do not exist in nature and select valuable mutant loci to improve crop quality.

10. Disease Resistance

Breeding stable disease-resistant plants is an economical and eco-friendly way to control crop diseases to sustain agricultural production. Rice blast is a major rice disease [111,112]. Mitigating or resolving rice blight is a problem for breeders to consider. Pi-d2 is an agriculturally important rice blast resistance gene. Previous studies have indicated that a single amino acid substitution at position 441 of the recessive allele of the Pi-d2 gene resulted in the loss of resistance to rice blast. CRISPR/Cas9 gene editing technology is powerless against single-base editing [113]. The use of an improved BE, hAID*Δ-XTEN-Cas9n-NLS (rBE5), was effective at rescuing resistance to rice blast in Pi-d2 mutants [114,115,116] (Table S1); rBE5 is a single-base editor for the introduction of human-derived AID cytosine deaminase. BEs can also be used in the treatment of other plant diseases. The OsSWEET14 gene is a susceptibility gene for rice leaf blight [117,118,119]. AFID systems (APOBEC-Cas9 fusion-induced deletion systems, AFIDs) are novel polynucleotide-targeted deletion systems in which wild-type SpCas9 is included with the cytosine deaminase APOBEC, uracil glycosylase (UDG), and purine-pyrimidine site-free cleavage enzymes (AP cleavage enzymes). The AFID system converts the cytidine on the non-target strand from APOBEC deaminase to uridine; UDG then excises the uracil from the uridine to create the AP site, which is removed by AP lyase; Cas9 cuts both strands to form the DSB, which leads to a “predictable” deletion from the deaminase C extension to the DSB via the NHEJ repair pathway. Researchers have used the AFID-3 system to target the effector binding element within the promoter of the OsSWEET14 gene in rice and obtained mutant plants in which polynucleotide sequences were deleted (Table S1); these mutants were subjected to leaf blight inoculation, and the results showed that, compared with plants in which only 1~2 bp were deleted, mutant plants in which polynucleotide sequences were deleted were more resistant to the leaf blight fungus [120]. With the development of sequencing technologies and bioinformatics, we can explore gene function more easily. More and more disease-resistant genes will be discovered, and using BE technology in combination with bioinformatics, precise editing of disease-resistant genes using BEs may be a promising approach to protect plants from biotic stresses [121].

11. Herbicide Resistance

Weeds compete with crop plants for nutrients, sunlight, and living space, disrupt airflow, intercept light, and promote pests and diseases in the field. Moreover, parasitic weeds absorb nutrients from crop plants, thereby decreasing the yield and quality of those crops [122,123]. Deploying weed management practices is therefore important [124,125,126].
BEs offer a precise and rapid way of generating new herbicide-resistant plant lines [127]. Acetyl lactate synthase (ALS) is a critical enzyme for synthesizing branched-chain amino acids and is an important target for herbicides, such as sulfonylureas and imidazolinones [104,128]. Studies have revealed that specific amino acid substitutions in the ALS gene can confer herbicide resistance to plants [129]. By using a CBE to target mutations in the base sequences corresponding to proline at position 171 (P171) and glycine at position 628 (G628) of the rice ALS gene, researchers obtained a series of ALS inhibitor-like herbicide resistance mutants (Figure 2b). Among them, P171S, P171A, P171Y, and P171F showed different levels of resistance after they were sprayed with five different types of ALS inhibitor-like herbicides [130,131] (Table S1). In addition, the triple-amino acid mutant P171F/G628E/G629S, in which there are mutations at the P171, G628, and G629 loci, showed high levels of resistance to all five ALS inhibitor-based herbicides [69]. Similarly, by altering specific base sites while aiming to maintain ALS activity, researchers are developing herbicide-tolerant lines of many plant species, including wheat [104], rice [69], maize [132], oilseed rape [62], tomato [133], watermelon [68], pear [134], and Arabidopsis [135] (Table S1). ACCase is a key enzyme in lipid biosynthesis. It represents the site of action for several commercially important herbicides, such as those of the aromatic phenoxy propionate (APP) class and the cyclohexanone (CHD) class [136]. Introducing a C2186R substitution in the rice ACCase gene via ABEs resulted in the production of haloxyfop-R-methyl-tolerant rice strains [137] (Table S1).
Other amino acid substitutions in ACCase confer resistance to rice haloxyfop, such as P1927F and W2125C, also identified through CRISPR-based screens [138,139]. In addition, editing PPO [140], EPSPS [141,142], TubA2 [127], and SF3B1 [143] has been reported to afford resistance to butafenacil, glyphosate, trifluralin, and herboxidiene (GEX1A) [144,145]. The advantage of base editing is that with targeted and precise substitution, more herbicide-resistant loci can be identified for agricultural production by constructing sgRNA libraries for gene screening (Figure 2a).

12. Multifunctional Single-Base Editors

Many important agricultural traits are associated with multiple heterogeneous base transformations [146,147]. To introduce both C-to-T and A-to-G substitutions in plant and mammalian cells, some teams have developed several new two-base editors by combining CBEs and ABEs. The two-base editors can efficiently generate mutations in two different bases simultaneously, enriching base editing tools and having important implications for species improvement and molecular evolution [148]. The team constructed a base editing system (GhBE3) for genetic transformation in cotton using cytosine deaminase (APOBEC1), Cas9 nickase (nCas9), and uracil glycosylase inhibitor (UGI). The system can efficiently and specifically introduce single-nucleotide mutations in cotton cells with high C–G to T–A single-base editing efficiency, with a C–T editing efficiency of 57.78%; this BE system will become a new and important tool for functional genomic research in cotton [61] (Table S1). In addition, the researchers constructed a novel saturation-targeted endogenous gene mutation BE, STEME; notably, the STEME double base editor can induce simultaneous C T and A G mutations at the target site with only one sgRNA guide, significantly increasing the saturation of base mutations in the target gene and the diversity of mutation types produced. By using STEME-1 and STEME-NG for targeted mutation of the rice OsACC gene, the researchers obtained an herbicide-resistant mutant [138] (Table S1). These two sets of STEMEs can accelerate trait development and function in any plant that can undergo CRISPR-based mutations. Based on the originally developed plant CT-CBE tool eCDAL, researchers constructed the double single-base editing vector pDuBE1 by fusing TadA-8e at the N-terminal end. pDuBE1 enabled double single-base editing at multiple loci in the rice genome. The application of pDuBE1 simultaneously induced point mutations in two herbicide-resistance-related genes, resulting in OsALS-P171F/OsACC1-I1781 V double mutant herbicide-resistant rice plants [148] (Table S1).
In addition, CBE can regulate gene expression by inserting premature stop codons at predetermined triplet codons via BEs by converting six codons—CAG, CGA, CAA, TCA, TAC, and TGG—into stop codons [149,150] (Figure 2d). Similarly, BE can also regulate gene expression by editing regulatory factors (Figure 2c). ACG or GTG may be substituted with ABEs for the initiation codon (ATG) [137,144,151]. This method of regulation of gene expression has been applied to a variety of plant species, including tomato, rice, wheat, and oilseed rape [144,152] (Table S1). Splicing eukaryotic mRNAs requires splice donor (GT) and splice acceptor (AG) sites, which can be disrupted by BEs. Therefore, by inducing misplacement, BEs can cause exon skipping [39], selective splicing, or intron retention. This method resulted in the generation of novel mutants in both Arabidopsis and rice [153]. We can expect that BE-mediated regulation of mRNA splicing and the introduction of the stop codon is also an approach to crop improvement [154].

13. Conclusions, Challenges, and Prospects

BEs can provide an effective and prospective genome editing approach for generating point mutations to control essential plant characteristics. With the creation of efficient multiple-base editing systems, efficient, precise, and targeted mutagenesis through genome editing sets the stage for the next generation of breeding strategies that will transform the future of agriculture [155]. In addition to the already discussed areas where base editing has been applied, base editing can also be used to edit organelle genomes [156,157,158,159,160]. The researchers split the DddAtox structural domain in the DddA protein into two parts, DddAtox-N and DddAtox-C, addressing its toxicity to mammalian cells, and added the mitochondrial guide peptide (MTS) gene. Following the successful development of an mtDNA editing system called DdCBE, the first molecular tool to enable precise editing of mtDNA, in 2022, researchers developed a new gene editing platform called transcription activator-like effector-linked dehydrogenase (TALED) [161,162,163]. TALED can perform A→G base conversion in mitochondria. The sterile lines needed to produce a good hybrid are cytoplasmic male sterility (CMS), which manifests as pollen abortive and is crossed with restorative lines to produce hybrid offspring [164]. Knockout of the orf79 gene in the mitochondria of the rice variety BTA and the orf125 gene in the mitochondria of the kale variety SW18 can restore fertility in sterile male lines using mitochondrial technology (mitochondria) [165]. The base editor is simple and precise and can be used to perform prescreens of laboratory experiments [161,162,163,165]. In addition, traditional breeding introduces resistance traits from wild species into cultivated species by crossbreeding, a method that takes years and is accompanied by unwanted mutations in traits other than the target trait [166,167]. Based on the understanding of the genetic and molecular laws of crop domestication, de novo domestication of naturally resistant wild plants using base editing techniques may be a novel strategy to obtain resistant crops [168,169].
However, several challenges remain to improve the efficiency of base editing and its applications in plant breeding. It should be pointed out that the past successes in plant breeding have been primarily built upon efficient exploitation of natural allelic variation at large numbers of loci existing in the crop germplasm resources. It remains a huge challenge to link different natural alleles at individual loci with phenotypic differences of target traits. This information is lacking and can help improve the efficiency of gene editing by narrowing specific genic region(s) for editing. Also, virtually all target traits, including yield potential, resistances or tolerances to abiotic and biotic stresses, and quality parameters, to be improved on by plant breeders involve large numbers of genes and complex gene networks. Thus, another challenge to improve the efficiency and effectiveness of base editing in plant breeding is how to accurately select target gene(s) to be edited, which depends on our knowledge of the genetic and molecular mechanisms underlying specific target traits and target genetic backgrounds to be improved. The greatest challenge, or question, is to what extent the gene editing technology applies to crops with large genomes, such as wheat and barley, because those crops have duplicated genomes or large numbers of duplicated genes, which make it even more difficult to select target genes and predict the consequences of products in addition to the transformation difficulty. Plant genome editing usually relies on Agrobacterium or gene gun-mediated introduction of sequence-specific nucleases such as CRISPR/Cas9 and selection marker expression frames into recipient plant cells, where the inserted exogenous genes can subsequently be isolated and removed from the chromosomes of the edited progeny plants by means of self or backcrossing. Agrobacterium transformation is more commonly used, but the process is time-consuming and laborious, especially for plants with complex ploidy, long breeding cycles, or asexual reproduction. Viral vector systems are ideal for the in vivo delivery and transient expression of exogenous genes and are an important complement to stable expression systems for transgenes [170]. However, there are still many important challenges to be overcome in the delivery system.
Technically, one constraint that limits base editing applications is the targeting scope of BEs, which rely on the PAM requirement of Cas proteins and the width of the catalytic reaction window [171,172,173]. To address this restriction, fusing deaminases with various Cas orthologous or engineered variants with altered or relaxed PAM specificities can be used to expand the editing scope, including SaCas9 [174], ScCas9 [175], SpCas9-NG [176], SpCas9-NRRH [177], SpCas9-NRTH [177] and so on. It’s worth noting that although using some variants has extended the range of base editors, it greatly reduces its editing efficiency and increases its dependence on target points. The SpRY variant is especially capable of targeting almost all PAMs, whereas the sgRNA editing frequencies showed extremely high results in low on-target editing frequencies. Therefore, further research is needed to improve the efficiency of base editors in maintaining the recognition of relaxed PAM.
In addition, whole-genome sequencing studies showed that BEs could generate genome-wide gRNA-dependent off-target mutations. Therefore, it is necessary to improve the editing specificity of ABEs and CBEs. Extending SgRNA guide sequences, using high-fidelity SpCas9 variants such as eSpCas9 [178], SpCas9-HF [179], and HypaCas9 [180], and delivering base editors as ribonucleoprotein complexes (RNPs) [181] were effective in reducing gRNA-dependent off-target effects. Besides, gRNA-independent off-target mutations were also found using CBEs but not ABEs in mice and rice. This may be due to the overexpression of deaminase, causing the whole genome, especially the gene enrichment region, to mutate randomly. Using an alternative deaminase to rAPOBEC1 or engineering the deaminase domain are effective strategies to reduce the off-target effects. For example, BE3 containing PpAPOBEC1, RrA3F, AmAPOBEC1, and SsAPOBEC3B were more specific than that containing rAPOBEC1 [182]. In rice cells, A3Bctd-VHM-BE3 and A3Bctd-KKR-BE3 exhibit markedly reduced gRNA-independent off-target editing. When editing target sites, BEs are often accompanied by unnecessary base substitutions [41,183]. One way to reduce the frequency of bystander mutations is to narrow the editing window. The width of the editing window is determined by the DNA base editor deaminase. Therefore, it takes lots of effort to modify cytidine deaminase by introducing amino acid mutations to narrow the editing window. YE1-BE3, YE2-BE3, EE-BE3 and YFE-BE4max based on
rAPOBEC1 with double or triple mutation reduced from a 5-nt editing window to 1–2 nts or 3 nts. However, to expand the application of BEs in plant genomes, base editors with expanded windows are also needed [25]. Therefore, an enriched toolbox of BEs developed from various Cas proteins linked with deaminases offers researchers numerous opportunities for functional genetic studies and crop breeding. While the combination of CBE and ABE can efficiently perform four base transitions (AT→GC, GC→AT), there is still a lack of more efficient editing tools (CGBEs are less efficient) for the other eight base transitions (AT→GC, GC→AT) and base insertion–deletions (indels)
Prime editing (PE) constitutes new precision gene editing tools developed by David Liu’s lab in 2019; the PE system consists of a Cas protein, reverse transcriptase, and pegRNA. pegRNA induces Cas9 nickase to cut the DNA strand and reverse transcriptase to synthesize a new DNA sequence from the unpaired pegRNA sequence, which is eventually integrated into the DNA to complete the gene editing. PEs can efficiently convert all 12 bases without relying on DSBs or donor DNA. They can also efficiently perform precise insertions (up to 44 bp) and deletions (up to 80 bp) of multiple bases [184]. PEs, therefore, constitute an all-purpose tool that enables major changes to the field of gene editing [185,186]. BEs allow single-base editing without creating DSBs, while the latest PEs not only induce all single-base mutations but also induce indels, and in this sense, PEs are more advanced than BEs [187]. However, PEs have more severe indel problems than do BEs. The detection of PE off-target sites still needs further confirmation. In addition, the safety of reverse transcriptase overexpressed in cells as a major building block remains an issue to be considered. Therefore, PE and BE functions complement each other to some extent. Ideally, we could first identify functional mutant loci by constructing a plant initiation editing library screen and then use BE for precise mutation targeting of the mutation. This integrated strategy would greatly facilitate direct gene evolution and germplasm innovation in crop improvement applications by targeting mutant loci through base editing. Gene editing technologies also present bottlenecks in crop improvement as current plant genetic transformation is inefficient; combining base editing techniques with traditional hybridization protocols to improve complex traits in crops is, therefore, a future trend [156,188]. It is hoped that, in the future, crops created by gene editing technology can be promoted for cultivation and create economic value in accordance with the policy on the regulation of gene editing technology and products.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/cimb45020059/s1, Table S1: Applications of base editing in plants.

Author Contributions

Y.L. and Y.J. wrote the first draft of the manuscript and organized the tables and figures; B.D., L.C., J.L. (Jing Liang) and J.Z. revised the manuscript; M.L. and J.L. (Juan Li) conceived and supervised the content of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (No. U19A2022 and No. 32000284), the Natural Science Foundation of Anhui Province (No. 2208085Y11, No. 2108085Y07), Science and Technology Major Projects of Anhui Province (No. 2021d06050002), the University Synergy Innovation Program of Anhui Province (GXXT-2021-056), Natural Science Research Project for Anhui Universities (KJ2021A0196) and Key Research and Development Project of Anhui Province (202004b11020009), the Improved Varieties Joint Research (Rice) Project of Anhui Province (the 14th five-year plan), and Key Research and Development Project of Anhui Province (202004b11020009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All of the data generated or analyzed during this study are included in this published article.

Acknowledgments

We thank Zhikang Li and Pengcheng Wei for providing valuable suggestions for the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gaj, T.; Gersbach, C.A.; Barbas, C.F., III. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Khalil, A.M. The genome editing revolution: Review. J. Genet. Eng. Biotechnol. 2020, 18, 68. [Google Scholar] [CrossRef] [PubMed]
  3. Gupta, D.; Bhattacharjee, O.; Mandal, D.; Sen, M.K.; Dey, D.; Dasgupta, A.; Kazi, T.A.; Gupta, R.; Sinharoy, S.; Acharya, K.; et al. CRISPR-Cas9 system: A new-fangled dawn in gene editing. Life Sci. 2019, 232, 116636. [Google Scholar] [CrossRef] [PubMed]
  4. Doench, J.G.; Fusi, N.; Sullender, M.; Hegde, M.; Vaimberg, E.W.; Donovan, K.F.; Smith, I.; Tothova, Z.; Wilen, C.; Orchard, R.; et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nat. Biotechnol. 2016, 34, 184–191. [Google Scholar] [CrossRef] [Green Version]
  5. Liu, G.; Lin, Q.; Jin, S.; Gao, C. The CRISPR-Cas toolbox and gene editing technologies. Mol. Cell 2022, 82, 333–347. [Google Scholar] [CrossRef]
  6. Chapman, J.R.; Taylor, M.R.; Boulton, S.J. Playing the end game: DNA double-strand break repair pathway choice. Mol. Cell 2012, 47, 497–510. [Google Scholar] [CrossRef] [Green Version]
  7. Kakarougkas, A.; Jeggo, P.A. DNA DSB repair pathway choice: An orchestrated handover mechanism. Br. J. Radiol. 2014, 87, 20130685. [Google Scholar] [CrossRef]
  8. Frit, P.; Ropars, V.; Modesti, M.; Charbonnier, J.B.; Calsou, P. Plugged into the Ku-DNA hub: The NHEJ network. Prog. Biophys. Mol. Biol. 2019, 147, 62–76. [Google Scholar] [CrossRef]
  9. Di Stazio, M.; Foschi, N.; Athanasakis, E.; Gasparini, P.; d’Adamo, A.P. Systematic analysis of factors that improve homologous direct repair (HDR) efficiency in CRISPR/Cas9 technique. PLoS ONE 2021, 16, e0247603. [Google Scholar] [CrossRef]
  10. Bennett, E.P.; Petersen, B.L.; Johansen, I.E.; Niu, Y.; Yang, Z.; Chamberlain, C.A.; Met, O.; Wandall, H.H.; Frodin, M. INDEL detection, the ‘Achilles heel’ of precise genome editing: A survey of methods for accurate profiling of gene editing induced indels. Nucleic Acids Res. 2020, 48, 11958–11981. [Google Scholar] [CrossRef]
  11. van de Kooij, B.; Kruswick, A.; van Attikum, H.; Yaffe, M.B. Multi-pathway DNA-repair reporters reveal competition between end-joining, single-strand annealing and homologous recombination at Cas9-induced DNA double-strand breaks. Nat. Commun. 2022, 13, 5295. [Google Scholar] [CrossRef] [PubMed]
  12. Scully, R.; Panday, A.; Elango, R.; Willis, N.A. DNA double-strand break repair-pathway choice in somatic mammalian cells. Nat. Rev. Mol. Cell Biol. 2019, 20, 698–714. [Google Scholar] [CrossRef] [PubMed]
  13. Schafer, K.A. The cell cycle: A review. Vet. Pathol. 1998, 35, 461–478. [Google Scholar] [CrossRef] [PubMed]
  14. Ravi, S.; Campagna, G.; Della Lucia, M.C.; Broccanello, C.; Bertoldo, G.; Chiodi, C.; Maretto, L.; Moro, M.; Eslami, A.S.; Srinivasan, S.; et al. SNP Alleles Associated With Low Bolting Tendency in Sugar Beet. Front. Plant Sci. 2021, 12, 693285. [Google Scholar] [CrossRef]
  15. McCarthy, J.J.; Hilfiker, R. The use of single-nucleotide polymorphism maps in pharmacogenomics. Nat. Biotechnol. 2000, 18, 505–508. [Google Scholar] [CrossRef]
  16. Sretenovic, S.; Liu, S.; Li, G.; Cheng, Y.; Fan, T.; Xu, Y.; Zhou, J.; Zheng, X.; Coleman, G.; Zhang, Y.; et al. Exploring C-To-G Base Editing in Rice, Tomato, and Poplar. Front. Genome Ed. 2021, 3, 756766. [Google Scholar] [CrossRef]
  17. Kantor, A.; McClements, M.E.; MacLaren, R.E. CRISPR-Cas9 DNA Base-Editing and Prime-Editing. Int. J. Mol. Sci. 2020, 21, 6240. [Google Scholar] [CrossRef] [PubMed]
  18. Thuronyi, B.W.; Koblan, L.W.; Levy, J.M.; Yeh, W.H.; Zheng, C.; Newby, G.A.; Wilson, C.; Bhaumik, M.; Shubina-Oleinik, O.; Holt, J.R.; et al. Continuous evolution of base editors with expanded target compatibility and improved activity. Nat. Biotechnol. 2019, 37, 1070–1079. [Google Scholar] [CrossRef] [PubMed]
  19. Negishi, K.; Kaya, H.; Abe, K.; Hara, N.; Saika, H.; Toki, S. An adenine base editor with expanded targeting scope using SpCas9-NGv1 in rice. Plant Biotechnol. J. 2019, 17, 1476–1478. [Google Scholar] [CrossRef]
  20. 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]
  21. Huang, T.P.; Newby, G.A.; Liu, D.R. Precision genome editing using cytosine and adenine base editors in mammalian cells. Nat. Protoc. 2021, 16, 1089–1128. [Google Scholar] [CrossRef] [PubMed]
  22. Matsoukas, I.G. Commentary: Programmable base editing of A.T to G.C in genomic DNA without DNA cleavage. Front. Genet. 2018, 9, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Komor, A.C.; Zhao, K.T.; Packer, M.S.; Gaudelli, N.M.; Waterbury, A.L.; Koblan, L.W.; Kim, Y.B.; Badran, A.H.; Liu, D.R. Improved base excision repair inhibition and bacteriophage Mu Gam protein yields C:G-to-T:A base editors with higher efficiency and product purity. Sci. Adv. 2017, 3, eaao4774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Koblan, L.W.; Doman, J.L.; Wilson, C.; Levy, J.M.; Tay, T.; Newby, G.A.; Maianti, J.P.; Raguram, A.; Liu, D.R. Improving cytidine and adenine base editors by expression optimization and ancestral reconstruction. Nat. Biotechnol. 2018, 36, 843–846. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, Y.B.; Komor, A.C.; Levy, J.M.; Packer, M.S.; Zhao, K.T.; Liu, D.R. Increasing the genome-targeting scope and precision of base editing with engineered Cas9-cytidine deaminase fusions. Nat. Biotechnol. 2017, 35, 371–376. [Google Scholar] [CrossRef]
  26. Olave, M.C.; Graham, R.P. Mismatch repair deficiency: The what, how and why it is important. Genes Chromosomes Cancer 2022, 61, 314–321. [Google Scholar] [CrossRef] [PubMed]
  27. Sameer, A.S.; Nissar, S.; Fatima, K. Mismatch repair pathway: Molecules, functions, and role in colorectal carcinogenesis. Eur. J. Cancer. Prev. 2014, 23, 246–257. [Google Scholar] [CrossRef]
  28. Villiger, L.; Rothgangl, T.; Witzigmann, D.; Oka, R.; Lin, P.J.C.; Qi, W.; Janjuha, S.; Berk, C.; Ringnalda, F.; Beattie, M.B.; et al. In vivo cytidine base editing of hepatocytes without detectable off-target mutations in RNA and DNA. Nat. Biomed. Eng. 2021, 5, 179–189. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, L.; Zhu, B.; Ru, G.; Meng, H.; Yan, Y.; Hong, M.; Zhang, D.; Luan, C.; Zhang, S.; Wu, H.; et al. Re-engineering the adenine deaminase TadA-8e for efficient and specific CRISPR-based cytosine base editing. Nat. Biotechnol. 2022, 10. [Google Scholar] [CrossRef]
  30. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable base editing of A*T to G*C in genomic DNA without DNA cleavage. Nature 2017, 551, 464–471. [Google Scholar] [CrossRef]
  31. Grunewald, J.; Zhou, R.; Garcia, S.P.; Iyer, S.; Lareau, C.A.; Aryee, M.J.; Joung, J.K. Transcriptome-wide off-target RNA editing induced by CRISPR-guided DNA base editors. Nature 2019, 569, 433–437. [Google Scholar] [CrossRef] [PubMed]
  32. Rees, H.A.; Wilson, C.; Doman, J.L.; Liu, D.R. Analysis and minimization of cellular RNA editing by DNA adenine base editors. Sci. Adv. 2019, 5, eaax5717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kim, H.S.; Jeong, Y.K.; Hur, J.K.; Kim, J.S.; Bae, S. Adenine base editors catalyze cytosine conversions in human cells. Nat. Biotechnol. 2019, 37, 1145–1148. [Google Scholar] [CrossRef] [PubMed]
  34. Montaldo, N.P.; Bordin, D.L.; Brambilla, A.; Rosinger, M.; Fordyce Martin, S.L.; Bjoras, K.O.; Bradamante, S.; Aas, P.A.; Furrer, A.; Olsen, L.C.; et al. Alkyladenine DNA glycosylase associates with transcription elongation to coordinate DNA repair with gene expression. Nat. Commun. 2019, 10, 5460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kurt, I.C.; Zhou, R.; Iyer, S.; Garcia, S.P.; Miller, B.R.; Langner, L.M.; Grunewald, J.; Joung, J.K. CRISPR C-to-G base editors for inducing targeted DNA transversions in human cells. Nat. Biotechnol. 2021, 39, 41–46. [Google Scholar] [CrossRef] [PubMed]
  36. Cortizas, E.M.; Zahn, A.; Safavi, S.; Reed, J.A.; Vega, F.; Di Noia, J.M.; Verdun, R.E. UNG protects B cells from AID-induced telomere loss. J. Exp. Med. 2016, 213, 2459–2472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cordoba-Canero, D.; Dubois, E.; Ariza, R.R.; Doutriaux, M.P.; Roldan-Arjona, T. Arabidopsis uracil DNA glycosylase (UNG) is required for base excision repair of uracil and increases plant sensitivity to 5-fluorouracil. J. Biol. Chem. 2010, 285, 7475–7483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Assefa, N.G.; Niiranen, L.; Johnson, K.A.; Leiros, H.K.; Smalas, A.O.; Willassen, N.P.; Moe, E. Structural and biophysical analysis of interactions between cod and human uracil-DNA N-glycosylase (UNG) and UNG inhibitor (Ugi). Acta Crystallogr. D Biol. Crystallogr. 2014, 70, 2093–2100. [Google Scholar] [CrossRef] [Green Version]
  39. Molla, K.A.; Yang, Y. CRISPR/Cas-Mediated Base Editing: Technical Considerations and Practical Applications. Trends Biotechnol. 2019, 37, 1121–1142. [Google Scholar] [CrossRef]
  40. Porto, E.M.; Komor, A.C.; Slaymaker, I.M.; Yeo, G.W. Base editing: Advances and therapeutic opportunities. Nat. Rev. Drug Discov. 2020, 19, 839–859. [Google Scholar] [CrossRef]
  41. Rees, H.A.; Liu, D.R. Base editing: Precision chemistry on the genome and transcriptome of living cells. Nat. Rev. Genet. 2018, 19, 770–788. [Google Scholar] [CrossRef] [PubMed]
  42. Allier, A.; Teyssedre, S.; Lehermeier, C.; Moreau, L.; Charcosset, A. Optimized breeding strategies to harness genetic resources with different performance levels. BMC Genom. 2020, 21, 349. [Google Scholar] [CrossRef] [PubMed]
  43. Till, B.J.; Reynolds, S.H.; Greene, E.A.; Codomo, C.A.; Enns, L.C.; Johnson, J.E.; Burtner, C.; Odden, A.R.; Young, K.; Taylor, N.E.; et al. Large-scale discovery of induced point mutations with high-throughput TILLING. Genome Res. 2003, 13, 524–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Qutub, M.; Chandran, S.; Rathinavel, K.; Sampathrajan, V.; Rajasekaran, R.; Manickam, S.; Adhimoolam, K.; Muniyandi, S.J.; Natesan, S. Improvement of a Yairipok Chujak Maize Landrace from North Eastern Himalayan Region for beta-Carotene Content through Molecular Marker-Assisted Backcross Breeding. Genes 2021, 12, 762. [Google Scholar] [CrossRef] [PubMed]
  45. Sserumaga, J.P.; Kayondo, S.I.; Kigozi, A.; Kiggundu, M.; Namazzi, C.; Walusimbi, K.; Bugeza, J.; Molly, A.; Mugerwa, S. Genome-wide diversity and structure variation among lablab [Lablab purpureus (L.) Sweet] accessions and their implication in a Forage breeding program. Genet. Resour. Crop Evol. 2021, 68, 2997–3010. [Google Scholar] [CrossRef]
  46. Hill, R.C.; Fast, B.J.; Herman, R.A. Transgenesis affects endogenous soybean allergen levels less than traditional breeding. Regul. Toxicol. Pharmacol. 2017, 89, 70–73. [Google Scholar] [CrossRef]
  47. Beans, C. Inner Workings: Crop researchers harness artificial intelligence to breed crops for the changing climate. Proc. Natl. Acad. Sci. USA 2020, 117, 27066–27069. [Google Scholar] [CrossRef]
  48. Kumar, K.; Gambhir, G.; Dass, A.; Tripathi, A.K.; Singh, A.; Jha, A.K.; Yadava, P.; Choudhary, M.; Rakshit, S. Genetically modified crops: Current status and future prospects. Planta 2020, 251, 91. [Google Scholar] [CrossRef]
  49. Kumlehn, J.; Pietralla, J.; Hensel, G.; Pacher, M.; Puchta, H. The CRISPR/Cas revolution continues: From efficient gene editing for crop breeding to plant synthetic biology. J. Integr. Plant Biol. 2018, 60, 1127–1153. [Google Scholar] [CrossRef]
  50. Schindele, A.; Dorn, A.; Puchta, H. CRISPR/Cas brings plant biology and breeding into the fast lane. Curr. Opin. Biotechnol. 2020, 61, 7–14. [Google Scholar] [CrossRef]
  51. Turner-Hissong, S.D.; Mabry, M.E.; Beissinger, T.M.; Ross-Ibarra, J.; Pires, J.C. Evolutionary insights into plant breeding. Curr. Opin. Plant Biol. 2020, 54, 93–100. [Google Scholar] [CrossRef] [PubMed]
  52. Saunders, T.L. The History of Transgenesis. Methods Mol. Biol. 2020, 2066, 1–26. [Google Scholar] [CrossRef] [PubMed]
  53. Araki, M.; Ishii, T. Towards social acceptance of plant breeding by genome editing. Trends Plant Sci. 2015, 20, 145–149. [Google Scholar] [CrossRef] [Green Version]
  54. Cubry, P.; Tranchant-Dubreuil, C.; Thuillet, A.C.; Monat, C.; Ndjiondjop, M.N.; Labadie, K.; Cruaud, C.; Engelen, S.; Scarcelli, N.; Rhone, B.; et al. The Rise and Fall of African Rice Cultivation Revealed by Analysis of 246 New Genomes. Curr. Biol. 2018, 28, 2274–2282 e2276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Eraslan, G.; Avsec, Z.; Gagneur, J.; Theis, F.J. Deep learning: New computational modelling techniques for genomics. Nat. Rev. Genet. 2019, 20, 389–403. [Google Scholar] [CrossRef]
  56. Hufford, M.B.; Seetharam, A.S.; Woodhouse, M.R.; Chougule, K.M.; Ou, S.; Liu, J.; Ricci, W.A.; Guo, T.; Olson, A.; Qiu, Y.; et al. De novo assembly, annotation, and comparative analysis of 26 diverse maize genomes. Science 2021, 373, 655–662. [Google Scholar] [CrossRef]
  57. International Wheat Genome Sequencing Consortium (IWGSC); Appels, R.; Eversole, K.; Stein, N.; Feuillet, C.; Keller, B.; Rogers, J.; Pozniak, C.J.; Choulet, F.; Distelfeld, A.; et al. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, eaar7191. [Google Scholar] [CrossRef] [Green Version]
  58. Earley, K.W.; Haag, J.R.; Pontes, O.; Opper, K.; Juehne, T.; Song, K.; Pikaard, C.S. Gateway-compatible vectors for plant functional genomics and proteomics. Plant J. 2006, 45, 616–629. [Google Scholar] [CrossRef]
  59. Doman, J.L.; Raguram, A.; Newby, G.A.; Liu, D.R. Evaluation and minimization of Cas9-independent off-target DNA editing by cytosine base editors. Nat. Biotechnol. 2020, 38, 620–628. [Google Scholar] [CrossRef]
  60. Li, J.; Sun, Y.; Du, J.; Zhao, Y.; Xia, L. Generation of Targeted Point Mutations in Rice by a Modified CRISPR/Cas9 System. Mol. Plant 2017, 10, 526–529. [Google Scholar] [CrossRef]
  61. Qin, L.; Li, J.; Wang, Q.; Xu, Z.; Sun, L.; Alariqi, M.; Manghwar, H.; Wang, G.; Li, B.; Ding, X.; et al. High-efficient and precise base editing of C*G to T*A in the allotetraploid cotton (Gossypium hirsutum) genome using a modified CRISPR/Cas9 system. Plant Biotechnol. J. 2020, 18, 45–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wu, J.; Chen, C.; Xian, G.; Liu, D.; Lin, L.; Yin, S.; Sun, Q.; Fang, Y.; Zhang, H.; Wang, Y. Engineering herbicide-resistant oilseed rape by CRISPR/Cas9-mediated cytosine base-editing. Plant Biotechnol. J. 2020, 18, 1857–1859. [Google Scholar] [CrossRef]
  63. Xu, R.; Liu, X.; Li, J.; Qin, R.; Wei, P. Identification of herbicide resistance OsACC1 mutations via in planta prime-editing-library screening in rice. Nat. Plants 2021, 7, 888–892. [Google Scholar] [CrossRef]
  64. Zong, Y.; Wang, Y.; Li, C.; Zhang, R.; Chen, K.; Ran, Y.; Qiu, J.L.; Wang, D.; Gao, C. Precise base editing in rice, wheat and maize with a Cas9-cytidine deaminase fusion. Nat. Biotechnol. 2017, 35, 438–440. [Google Scholar] [CrossRef] [PubMed]
  65. Yan, D.; Ren, B.; Liu, L.; Yan, F.; Li, S.; Wang, G.; Sun, W.; Zhou, X.; Zhou, H. High-efficiency and multiplex adenine base editing in plants using new TadA variants. Mol. Plant 2021, 14, 722–731. [Google Scholar] [CrossRef] [PubMed]
  66. Hunziker, J.; Nishida, K.; Kondo, A.; Kishimoto, S.; Ariizumi, T.; Ezura, H. Multiple gene substitution by Target-AID base-editing technology in tomato. Sci. Rep. 2020, 10, 20471. [Google Scholar] [CrossRef]
  67. Xing, S.; Chen, K.; Zhu, H.; Zhang, R.; Zhang, H.; Li, B.; Gao, C. Fine-tuning sugar content in strawberry. Genome Biol. 2020, 21, 230. [Google Scholar] [CrossRef]
  68. Tian, S.; Jiang, L.; Cui, X.; Zhang, J.; Guo, S.; Li, M.; Zhang, H.; Ren, Y.; Gong, G.; Zong, M.; et al. Engineering herbicide-resistant watermelon variety through CRISPR/Cas9-mediated base-editing. Plant Cell Rep. 2018, 37, 1353–1356. [Google Scholar] [CrossRef]
  69. Zhang, R.; Chen, S.; Meng, X.; Chai, Z.; Wang, D.; Yuan, Y.; Chen, K.; Jiang, L.; Li, J.; Gao, C. Generating broad-spectrum tolerance to ALS-inhibiting herbicides in rice by base editing. Sci. China Life Sci. 2021, 64, 1624–1633. [Google Scholar] [CrossRef]
  70. Breseghello, F.; Coelho, A.S. Traditional and modern plant breeding methods with examples in rice (Oryza sativa L.). J. Agric. Food Chem. 2013, 61, 8277–8286. [Google Scholar] [CrossRef]
  71. Klompe, S.E.; Vo, P.L.H.; Halpin-Healy, T.S.; Sternberg, S.H. Transposon-encoded CRISPR-Cas systems direct RNA-guided DNA integration. Nature 2019, 571, 219–225. [Google Scholar] [CrossRef] [PubMed]
  72. Strecker, J.; Ladha, A.; Gardner, Z.; Schmid-Burgk, J.L.; Makarova, K.S.; Koonin, E.V.; Zhang, F. RNA-guided DNA insertion with CRISPR-associated transposases. Science 2019, 365, 48–53. [Google Scholar] [CrossRef] [PubMed]
  73. Tilman, D.; Balzer, C.; Hill, J.; Befort, B.L. Global food demand and the sustainable intensification of agriculture. Proc. Natl. Acad. Sci. USA 2011, 108, 20260–20264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Mazza, J.J. Climate Change and Agriculture: Future Implications. Wis. Med. J. 2017, 116, 191. [Google Scholar]
  75. Hua, K.; Tao, X.; Zhu, J.K. Expanding the base editing scope in rice by using Cas9 variants. Plant Biotechnol. J. 2019, 17, 499–504. [Google Scholar] [CrossRef] [Green Version]
  76. Hua, K.; Tao, X.; Han, P.; Wang, R.; Zhu, J.K. Genome Engineering in Rice Using Cas9 Variants that Recognize NG PAM Sequences. Mol. Plant 2019, 12, 1003–1014. [Google Scholar] [CrossRef]
  77. Ren, L.; Tang, D.; Zhao, T.; Zhang, F.; Liu, C.; Xue, Z.; Shi, W.; Du, G.; Shen, Y.; Li, Y.; et al. OsSPL regulates meiotic fate acquisition in rice. New Phytol. 2018, 218, 789–803. [Google Scholar] [CrossRef] [Green Version]
  78. Miura, K.; Ikeda, M.; Matsubara, A.; Song, X.J.; Ito, M.; Asano, K.; Matsuoka, M.; Kitano, H.; Ashikari, M. OsSPL14 promotes panicle branching and higher grain productivity in rice. Nat. Genet. 2010, 42, 545–549. [Google Scholar] [CrossRef]
  79. Jiao, Y.; Wang, Y.; Xue, D.; Wang, J.; Yan, M.; Liu, G.; Dong, G.; Zeng, D.; Lu, Z.; Zhu, X.; et al. Regulation of OsSPL14 by OsmiR156 defines ideal plant architecture in rice. Nat. Genet. 2010, 42, 541–544. [Google Scholar] [CrossRef]
  80. Ren, Q.; Sretenovic, S.; Liu, G.; Zhong, Z.; Wang, J.; Huang, L.; Tang, X.; Guo, Y.; Liu, L.; Wu, Y.; et al. Improved plant cytosine base editors with high editing activity, purity, and specificity. Plant Biotechnol. J. 2021, 19, 2052–2068. [Google Scholar] [CrossRef]
  81. Hua, K.; Tao, X.; Yuan, F.; Wang, D.; Zhu, J.K. Precise A.T to G.C Base Editing in the Rice Genome. Mol. Plant 2018, 11, 627–630. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, J.; Xin, X.; He, Y.; Chen, H.; Li, Q.; Tang, X.; Zhong, Z.; Deng, K.; Zheng, X.; Akher, S.A.; et al. Multiplex QTL editing of grain-related genes improves yield in elite rice varieties. Plant Cell Rep. 2019, 38, 475–485. [Google Scholar] [CrossRef] [PubMed]
  83. Yang, Z.; Bai, Z.; Li, X.; Wang, P.; Wu, Q.; Yang, L.; Li, L.; Li, X. SNP identification and allelic-specific PCR markers development for TaGW2, a gene linked to wheat kernel weight. Theor. Appl. Genet. 2012, 125, 1057–1068. [Google Scholar] [CrossRef] [PubMed]
  84. Ashikari, M.; Sakakibara, H.; Lin, S.; Yamamoto, T.; Takashi, T.; Nishimura, A.; Angeles, E.R.; Qian, Q.; Kitano, H.; Matsuoka, M. Cytokinin oxidase regulates rice grain production. Science 2005, 309, 741–745. [Google Scholar] [CrossRef] [PubMed]
  85. Gaenssle, A.L.O.; van der Maarel, M.; Jurak, E. The influence of amylose content on the modification of starches by glycogen branching enzymes. Food Chem. 2022, 393, 133294. [Google Scholar] [CrossRef]
  86. Fasahat, P.; Rahman, S.; Ratnam, W. Genetic controls on starch amylose content in wheat and rice grains. J. Genet. 2014, 93, 279–292. [Google Scholar] [CrossRef]
  87. Wang, W.; Mauleon, R.; Hu, Z.; Chebotarov, D.; Tai, S.; Wu, Z.; Li, M.; Zheng, T.; Fuentes, R.R.; Zhang, F.; et al. Genomic variation in 3,010 diverse accessions of Asian cultivated rice. Nature 2018, 557, 43–49. [Google Scholar] [CrossRef] [Green Version]
  88. Tetlow, I.J.; Emes, M.J. A review of starch-branching enzymes and their role in amylopectin biosynthesis. IUBMB Life 2014, 66, 546–558. [Google Scholar] [CrossRef]
  89. Zhong, H.; Liu, S.; Zhao, G.; Zhang, C.; Peng, Z.; Wang, Z.; Yang, J.; Li, Y. Genetic Diversity Relationship Between Grain Quality and Appearance in Rice. Front. Plant Sci. 2021, 12, 708996. [Google Scholar] [CrossRef]
  90. Keeratiburana, T.; Hansen, A.R.; Soontaranon, S.; Blennow, A.; Tongta, S. Porous high amylose rice starch modified by amyloglucosidase and maltogenic alpha-amylase. Carbohydr. Polym. 2020, 230, 115611. [Google Scholar] [CrossRef]
  91. Huang, L.; Li, Q.; Zhang, C.; Chu, R.; Gu, Z.; Tan, H.; Zhao, D.; Fan, X.; Liu, Q. Creating novel Wx alleles with fine-tuned amylose levels and improved grain quality in rice by promoter editing using CRISPR/Cas9 system. Plant Biotechnol. J. 2020, 18, 2164–2166. [Google Scholar] [CrossRef] [PubMed]
  92. Xu, Y.; Lin, Q.; Li, X.; Wang, F.; Chen, Z.; Wang, J.; Li, W.; Fan, F.; Tao, Y.; Jiang, Y.; et al. Fine-tuning the amylose content of rice by precise base editing of the Wx gene. Plant Biotechnol. J. 2021, 19, 11–13. [Google Scholar] [CrossRef] [PubMed]
  93. Adegoke, T.V.; Wang, Y.; Chen, L.; Wang, H.; Liu, W.; Liu, X.; Cheng, Y.C.; Tong, X.; Ying, J.; Zhang, J. Posttranslational Modification of Waxy to Genetically Improve Starch Quality in Rice Grain. Int. J. Mol. Sci. 2021, 22, 4845. [Google Scholar] [CrossRef] [PubMed]
  94. Monsur, M.B.; Ni, C.; Xiangjin, W.; Lihong, X.; Guiai, J.; Shaoqing, T.; Sreenivasulu, N.; Gaoneng, S.; Peisong, H. Improved Eating and Cooking Quality of indica Rice Cultivar YK17 via Adenine Base Editing of Wx Allele of Granule-Bound Starch Synthase I (GBSS I). Rice Science 2021, 28, 427–430. [Google Scholar] [CrossRef]
  95. Tian, Z.; Qian, Q.; Liu, Q.; Yan, M.; Liu, X.; Yan, C.; Liu, G.; Gao, Z.; Tang, S.; Zeng, D.; et al. Allelic diversities in rice starch biosynthesis lead to a diverse array of rice eating and cooking qualities. Proc. Natl. Acad. Sci. USA 2009, 106, 21760–21765. [Google Scholar] [CrossRef] [Green Version]
  96. Nakamura, Y. Towards a better understanding of the metabolic system for amylopectin biosynthesis in plants: Rice endosperm as a model tissue. Plant Cell Physiol. 2002, 43, 718–725. [Google Scholar] [CrossRef] [Green Version]
  97. Nishi, A.; Nakamura, Y.; Tanaka, N.; Satoh, H. Biochemical and genetic analysis of the effects of amylose-extender mutation in rice endosperm. Plant Physiol. 2001, 127, 459–472. [Google Scholar] [CrossRef]
  98. Cruz Bojorquez, R.M.; Gonzalez Gallego, J.; Sanchez Collado, P. Functional properties and health benefits of lycopene. Nutr. Hosp. 2013, 28, 6–15. [Google Scholar] [CrossRef]
  99. Zhu, R.; Chen, B.; Bai, Y.; Miao, T.; Rui, L.; Zhang, H.; Xia, B.; Li, Y.; Gao, S.; Wang, X.D.; et al. Lycopene in protection against obesity and diabetes: A mechanistic review. Pharmacol. Res. 2020, 159, 104966. [Google Scholar] [CrossRef]
  100. Li, N.; Wu, X.; Zhuang, W.; Xia, L.; Chen, Y.; Wu, C.; Rao, Z.; Du, L.; Zhao, R.; Yi, M.; et al. Tomato and lycopene and multiple health outcomes: Umbrella review. Food Chem. 2021, 343, 128396. [Google Scholar] [CrossRef]
  101. Nishida, K.; Arazoe, T.; Yachie, N.; Banno, S.; Kakimoto, M.; Tabata, M.; Mochizuki, M.; Miyabe, A.; Araki, M.; Hara, K.Y.; et al. Targeted nucleotide editing using hybrid prokaryotic and vertebrate adaptive immune systems. Science 2016, 353, aaf8729. [Google Scholar] [CrossRef] [PubMed]
  102. Khan, U.M.; Sevindik, M.; Zarrabi, A.; Nami, M.; Ozdemir, B.; Kaplan, D.N.; Selamoglu, Z.; Hasan, M.; Kumar, M.; Alshehri, M.M.; et al. Lycopene: Food Sources, Biological Activities, and Human Health Benefits. Oxid. Med. Cell Longev. 2021, 2021, 2713511. [Google Scholar] [CrossRef]
  103. Hunziker, J.; Nishida, K.; Kondo, A.; Ariizumi, T.; Ezura, H. Phenotypic Characterization of High Carotenoid Tomato Mutants Generated by the Target-AID Base-Editing Technology. Front. Plant Sci. 2022, 13, 848560. [Google Scholar] [CrossRef] [PubMed]
  104. Zong, Y.; Song, Q.; Li, C.; Jin, S.; Zhang, D.; Wang, Y.; Qiu, J.L.; Gao, C. Efficient C-to-T base editing in plants using a fusion of nCas9 and human APOBEC3A. Nat. Biotechnol. 2018, 36, 950–953. [Google Scholar] [CrossRef]
  105. Zhang, J.; Liu, Y.X.; Zhang, N.; Hu, B.; Jin, T.; Xu, H.; Qin, Y.; Yan, P.; Zhang, X.; Guo, X.; et al. NRT1.1B is associated with root microbiota composition and nitrogen use in field-grown rice. Nat. Biotechnol. 2019, 37, 676–684. [Google Scholar] [CrossRef] [PubMed]
  106. Hu, B.; Wang, W.; Ou, S.; Tang, J.; Li, H.; Che, R.; Zhang, Z.; Chai, X.; Wang, H.; Wang, Y.; et al. Variation in NRT1.1B contributes to nitrate-use divergence between rice subspecies. Nat. Genet. 2015, 47, 834–838. [Google Scholar] [CrossRef]
  107. Asano, K.; Hirano, K.; Ueguchi-Tanaka, M.; Angeles-Shim, R.B.; Komura, T.; Satoh, H.; Kitano, H.; Matsuoka, M.; Ashikari, M. Isolation and characterization of dominant dwarf mutants, Slr1-d, in rice. Mol. Genet. Genom. 2009, 281, 223–231. [Google Scholar] [CrossRef]
  108. Ikeda, A.; Ueguchi-Tanaka, M.; Sonoda, Y.; Kitano, H.; Koshioka, M.; Futsuhara, Y.; Matsuoka, M.; Yamaguchi, J. slender rice, a constitutive gibberellin response mutant, is caused by a null mutation of the SLR1 gene, an ortholog of the height-regulating gene GAI/RGA/RHT/D8. Plant Cell 2001, 13, 999–1010. [Google Scholar] [CrossRef] [Green Version]
  109. Lu, Y.; Zhu, J.K. Precise Editing of a Target Base in the Rice Genome Using a Modified CRISPR/Cas9 System. Mol. Plant 2017, 10, 523–525. [Google Scholar] [CrossRef] [Green Version]
  110. Cheng, H.; Hao, M.; Ding, B.; Mei, D.; Wang, W.; Wang, H.; Zhou, R.; Liu, J.; Li, C.; Hu, Q. Base editing with high efficiency in allotetraploid oilseed rape by A3A-PBE system. Plant Biotechnol. J. 2021, 19, 87–97. [Google Scholar] [CrossRef]
  111. Miah, G.; Rafii, M.Y.; Ismail, M.R.; Puteh, A.B.; Rahim, H.A.; Asfaliza, R.; Latif, M.A. Blast resistance in rice: A review of conventional breeding to molecular approaches. Mol. Biol. Rep. 2013, 40, 2369–2388. [Google Scholar] [CrossRef] [PubMed]
  112. Valent, B. The Impact of Blast Disease: Past, Present, and Future. Methods Mol. Biol. 2021, 2356, 1–18. [Google Scholar] [CrossRef] [PubMed]
  113. Kouzai, Y.; Kaku, H.; Shibuya, N.; Minami, E.; Nishizawa, Y. Expression of the chimeric receptor between the chitin elicitor receptor CEBiP and the receptor-like protein kinase Pi-d2 leads to enhanced responses to the chitin elicitor and disease resistance against Magnaporthe oryzae in rice. Plant Mol. Biol. 2013, 81, 287–295. [Google Scholar] [CrossRef]
  114. Ren, B.; Yan, F.; Kuang, Y.; Li, N.; Zhang, D.; Zhou, X.; Lin, H.; Zhou, H. Improved Base Editor for Efficiently Inducing Genetic Variations in Rice with CRISPR/Cas9-Guided Hyperactive hAID Mutant. Mol. Plant 2018, 11, 623–626. [Google Scholar] [CrossRef] [Green Version]
  115. Li, J.B.; Sun, Y.D.; Liu, H.; Wang, Y.Y.; Jia, Y.L.; Xu, M.H. Natural variation of rice blast resistance gene Pi-d2. Genet. Mol. Res. 2015, 14, 1235–1249. [Google Scholar] [CrossRef] [PubMed]
  116. Chen, X.; Shang, J.; Chen, D.; Lei, C.; Zou, Y.; Zhai, W.; Liu, G.; Xu, J.; Ling, Z.; Cao, G.; et al. A B-lectin receptor kinase gene conferring rice blast resistance. Plant J. 2006, 46, 794–804. [Google Scholar] [CrossRef]
  117. Chen, L.Q.; Hou, B.H.; Lalonde, S.; Takanaga, H.; Hartung, M.L.; Qu, X.Q.; Guo, W.J.; Kim, J.G.; Underwood, W.; Chaudhuri, B.; et al. Sugar transporters for intercellular exchange and nutrition of pathogens. Nature 2010, 468, 527–532. [Google Scholar] [CrossRef] [Green Version]
  118. White, F.F.; Potnis, N.; Jones, J.B.; Koebnik, R. The type III effectors of Xanthomonas. Mol. Plant Pathol. 2009, 10, 749–766. [Google Scholar] [CrossRef]
  119. Streubel, J.; Pesce, C.; Hutin, M.; Koebnik, R.; Boch, J.; Szurek, B. Five phylogenetically close rice SWEET genes confer TAL effector-mediated susceptibility to Xanthomonas oryzae pv. oryzae. New Phytol. 2013, 200, 808–819. [Google Scholar] [CrossRef]
  120. Wang, S.; Zong, Y.; Lin, Q.; Zhang, H.; Chai, Z.; Zhang, D.; Chen, K.; Qiu, J.L.; Gao, C. Precise, predictable multi-nucleotide deletions in rice and wheat using APOBEC-Cas9. Nat. Biotechnol. 2020, 38, 1460–1465. [Google Scholar] [CrossRef]
  121. Bastet, A.; Zafirov, D.; Giovinazzo, N.; Guyon-Debast, A.; Nogue, F.; Robaglia, C.; Gallois, J.L. Mimicking natural polymorphism in eIF4E by CRISPR-Cas9 base editing is associated with resistance to potyviruses. Plant Biotechnol. J. 2019, 17, 1736–1750. [Google Scholar] [CrossRef] [PubMed]
  122. Scarrow, R. Weeds represent growing threat to crop yields. Nat. Plants 2022, 8, 7. [Google Scholar] [CrossRef] [PubMed]
  123. Edwards, R.; Hannah, M. Focus on weed control. Plant Physiol. 2014, 166, 1087–1089. [Google Scholar] [CrossRef] [Green Version]
  124. Bernasconi, P.; Woodworth, A.R.; Rosen, B.A.; Subramanian, M.V.; Siehl, D.L. A naturally occurring point mutation confers broad range tolerance to herbicides that target acetolactate synthase. J. Biol. Chem. 1996, 271, 13925. [Google Scholar] [CrossRef]
  125. Kaur, R.; Kaur, S.; Deol, J.S.; Sharma, R.; Kaur, T.; Brar, A.S.; Choudhary, O.P. Soil Properties and Weed Dynamics in Wheat as Affected by Rice Residue Management in the Rice-Wheat Cropping System in South Asia: A Review. Plants 2021, 10, 953. [Google Scholar] [CrossRef]
  126. Oliveira, M.C.; Osipitan, O.A.; Begcy, K.; Werle, R. Cover crops, hormones and herbicides: Priming an integrated weed management strategy. Plant Sci. 2020, 301, 110550. [Google Scholar] [CrossRef]
  127. Liu, L.; Kuang, Y.; Yan, F.; Li, S.; Ren, B.; Gosavi, G.; Spetz, C.; Li, X.; Wang, X.; Zhou, X.; et al. Developing a novel artificial rice germplasm for dinitroaniline herbicide resistance by base editing of OsTubA2. Plant Biotechnol. J. 2021, 19, 5–7. [Google Scholar] [CrossRef] [PubMed]
  128. Yu, Q.; Powles, S.B. Resistance to AHAS inhibitor herbicides: Current understanding. Pest Manag. Sci. 2014, 70, 1340–1350. [Google Scholar] [CrossRef]
  129. Liu, W.; Bai, S.; Jia, S.; Guo, W.; Zhang, L.; Li, W.; Wang, J. Comparison of ALS functionality and plant growth in ALS-inhibitor susceptible and resistant Myosoton aquaticum L. Pestic. Biochem. Physiol. 2017, 142, 111–116. [Google Scholar] [CrossRef]
  130. Hussain, A.; Ding, X.; Alariqi, M.; Manghwar, H.; Hui, F.; Li, Y.; Cheng, J.; Wu, C.; Cao, J.; Jin, S. Herbicide Resistance: Another Hot Agronomic Trait for Plant Genome Editing. Plants 2021, 10, 621. [Google Scholar] [CrossRef]
  131. Palmieri, V.E.; Alvarez, C.E.; Permingeat, H.R.; Perotti, V.E. A122S, A205V, D376E, W574L and S653N substitutions in acetolactate synthase (ALS) from Amaranthus palmeri show different functional impacts on herbicide resistance. Pest Manag. Sci. 2022, 78, 749–757. [Google Scholar] [CrossRef]
  132. Li, Y.; Zhu, J.; Wu, H.; Liu, C.; Huang, C.; Lan, J.; Zhao, Y.; Xie, C. Precise base editing of non-allelic acetolactate synthase genes confers sulfonylurea herbicide resistance in maize. Crop J. 2020, 8, 446–456. [Google Scholar] [CrossRef]
  133. Veillet, F.; Perrot, L.; Chauvin, L.; Kermarrec, M.P.; Guyon-Debast, A.; Chauvin, J.E.; Nogue, F.; Mazier, M. Transgene-Free Genome Editing in Tomato and Potato Plants Using Agrobacterium-Mediated Delivery of a CRISPR/Cas9 Cytidine Base Editor. Int. J. Mol. Sci. 2019, 20, 402. [Google Scholar] [CrossRef] [Green Version]
  134. Malabarba, J.; Chevreau, E.; Dousset, N.; Veillet, F.; Moizan, J.; Vergne, E. New Strategies to Overcome Present CRISPR/Cas9 Limitations in Apple and Pear: Efficient Dechimerization and Base Editing. Int. J. Mol. Sci. 2020, 22, 319. [Google Scholar] [CrossRef] [PubMed]
  135. Kang, B.C.; Yun, J.Y.; Kim, S.T.; Shin, Y.; Ryu, J.; Choi, M.; Woo, J.W.; Kim, J.S. Precision genome engineering through adenine base editing in plants. Nat. Plants 2018, 4, 427–431. [Google Scholar] [CrossRef] [PubMed]
  136. Vazquez-Garcia, J.G.; Alcantara-de la Cruz, R.; Palma-Bautista, C.; Rojano-Delgado, A.M.; Cruz-Hipolito, H.E.; Torra, J.; Barro, F.; De Prado, R. Accumulation of Target Gene Mutations Confers Multiple Resistance to ALS, ACCase, and EPSPS Inhibitors in Lolium Species in Chile. Front. Plant Sci. 2020, 11, 553948. [Google Scholar] [CrossRef] [PubMed]
  137. Li, C.; Zong, Y.; Wang, Y.; Jin, S.; Zhang, D.; Song, Q.; Zhang, R.; Gao, C. Expanded base editing in rice and wheat using a Cas9-adenosine deaminase fusion. Genome Biol. 2018, 19, 59. [Google Scholar] [CrossRef] [Green Version]
  138. Li, C.; Zhang, R.; Meng, X.; Chen, S.; Zong, Y.; Lu, C.; Qiu, J.L.; Chen, Y.H.; Li, J.; Gao, C. Targeted, random mutagenesis of plant genes with dual cytosine and adenine base editors. Nat. Biotechnol. 2020, 38, 875–882. [Google Scholar] [CrossRef]
  139. Liu, X.; Qin, R.; Li, J.; Liao, S.; Shan, T.; Xu, R.; Wu, D.; Wei, P. A CRISPR-Cas9-mediated domain-specific base-editing screen enables functional assessment of ACCase variants in rice. Plant Biotechnol. J. 2020, 18, 1845–1847. [Google Scholar] [CrossRef] [Green Version]
  140. de Pater, S.; Klemann, B.; Hooykaas, P.J.J. True gene-targeting events by CRISPR/Cas-induced DSB repair of the PPO locus with an ectopically integrated repair template. Sci. Rep. 2018, 8, 3338. [Google Scholar] [CrossRef] [Green Version]
  141. Hummel, A.W.; Chauhan, R.D.; Cermak, T.; Mutka, A.M.; Vijayaraghavan, A.; Boyher, A.; Starker, C.G.; Bart, R.; Voytas, D.F.; Taylor, N.J. Allele exchange at the EPSPS locus confers glyphosate tolerance in cassava. Plant Biotechnol. J. 2018, 16, 1275–1282. [Google Scholar] [CrossRef]
  142. Endo, M.; Mikami, M.; Endo, A.; Kaya, H.; Itoh, T.; Nishimasu, H.; Nureki, O.; Toki, S. Genome editing in plants by engineered CRISPR-Cas9 recognizing NG PAM. Nat. Plants 2019, 5, 14–17. [Google Scholar] [CrossRef]
  143. Butt, H.; Eid, A.; Momin, A.A.; Bazin, J.; Crespi, M.; Arold, S.T.; Mahfouz, M.M. CRISPR directed evolution of the spliceosome for resistance to splicing inhibitors. Genome Biol. 2019, 20, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Shimatani, Z.; Kashojiya, S.; Takayama, M.; Terada, R.; Arazoe, T.; Ishii, H.; Teramura, H.; Yamamoto, T.; Komatsu, H.; Miura, K.; et al. Targeted base editing in rice and tomato using a CRISPR-Cas9 cytidine deaminase fusion. Nat. Biotechnol. 2017, 35, 441–443. [Google Scholar] [CrossRef] [PubMed]
  145. Zhu, H.; Li, C.; Gao, C. Applications of CRISPR-Cas in agriculture and plant biotechnology. Nat. Rev. Mol. Cell Biol. 2020, 21, 661–677. [Google Scholar] [CrossRef] [PubMed]
  146. Ganal, M.W.; Altmann, T.; Roder, M.S. SNP identification in crop plants. Curr. Opin. Plant Biol. 2009, 12, 211–217. [Google Scholar] [CrossRef]
  147. Mafra, G.S.; Amaral Junior, A.T.D.; Almeida Filho, J.E.; Vivas, M.; Santos, P.; Santos, J.S.; Pena, G.F.; Lima, V.J.; Kamphorst, S.H.; Oliveira, F.T.; et al. SNP-based mixed model association of growth- and yield-related traits in popcorn. PLoS ONE 2019, 14, e0218552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Xu, R.; Kong, F.; Qin, R.; Li, J.; Liu, X.; Wei, P. Development of an efficient plant dual cytosine and adenine editor. J. Integr. Plant Biol. 2021, 63, 1600–1605. [Google Scholar] [CrossRef]
  149. Chemla, Y.; Ozer, E.; Algov, I.; Alfonta, L. Context effects of genetic code expansion by stop codon suppression. Curr. Opin. Chem. Biol. 2018, 46, 146–155. [Google Scholar] [CrossRef]
  150. Billon, P.; Bryant, E.E.; Joseph, S.A.; Nambiar, T.S.; Hayward, S.B.; Rothstein, R.; Ciccia, A. CRISPR-Mediated Base Editing Enables Efficient Disruption of Eukaryotic Genes through Induction of STOP Codons. Mol. Cell 2017, 67, 1068–1079 e1064. [Google Scholar] [CrossRef] [Green Version]
  151. Wang, X.; Liu, Z.; Li, G.; Dang, L.; Huang, S.; He, L.; Ma, Y.; Li, C.; Liu, M.; Yang, G.; et al. Efficient Gene Silencing by Adenine Base Editor-Mediated Start Codon Mutation. Mol. Ther. 2020, 28, 431–440. [Google Scholar] [CrossRef]
  152. Komatsu, A.; Ohtake, M.; Shimatani, Z.; Nishida, K. Production of Herbicide-Sensitive Strain to Prevent Volunteer Rice Infestation Using a CRISPR-Cas9 Cytidine Deaminase Fusion. Front. Plant Sci. 2020, 11, 925. [Google Scholar] [CrossRef]
  153. Li, Z.; Xiong, X.; Wang, F.; Liang, J.; Li, J.F. Gene disruption through base editing-induced messenger RNA missplicing in plants. New Phytol. 2019, 222, 1139–1148. [Google Scholar] [CrossRef] [PubMed]
  154. Ichinose, M.; Sugita, M. RNA Editing and Its Molecular Mechanism in Plant Organelles. Genes 2016, 8, 5. [Google Scholar] [CrossRef] [Green Version]
  155. Chen, Y.; Cheng, M.; Li, Y.; Wang, L.; Fang, L.; Cao, Y.; Song, H. Highly efficient multiplex base editing: One-shot deactivation of eight genes in Shewanella oneidensis MR-1. Synth. Syst. Biotechnol. 2023, 8, 1–10. [Google Scholar] [CrossRef]
  156. Yarra, R.; Sahoo, L. Base editing in rice: Current progress, advances, limitations, and future perspectives. Plant Cell Rep. 2021, 40, 595–604. [Google Scholar] [CrossRef] [PubMed]
  157. Haroon, M.; Wang, X.; Afzal, R.; Zafar, M.M.; Idrees, F.; Batool, M.; Khan, A.S.; Imran, M. Novel Plant Breeding Techniques Shake Hands with Cereals to Increase Production. Plants 2022, 11, 1052. [Google Scholar] [CrossRef] [PubMed]
  158. Fiaz, S.; Ahmar, S.; Saeed, S.; Riaz, A.; Mora-Poblete, F.; Jung, K.H. Evolution and Application of Genome Editing Techniques for Achieving Food and Nutritional Security. Int. J. Mol. Sci. 2021, 22, 5585. [Google Scholar] [CrossRef]
  159. Bacman, S.R.; Moraes, C.T. Mitochondrial DNA Base Editing: Good Editing Things Still Come in Small Packages. Mol. Cell 2020, 79, 708–709. [Google Scholar] [CrossRef]
  160. Verechshagina, N.; Nikitchina, N.; Yamada, Y.; Harashima capital En, C.; Tanaka, M.; Orishchenko, K.; Mazunin, I. Future of human mitochondrial DNA editing technologies. Mitochondrial. DNA. A. DNA. Mapp. Seq. Anal. 2019, 30, 214–221. [Google Scholar] [CrossRef]
  161. Mok, B.Y.; de Moraes, M.H.; Zeng, J.; Bosch, D.E.; Kotrys, A.V.; Raguram, A.; Hsu, F.; Radey, M.C.; Peterson, S.B.; Mootha, V.K.; et al. A bacterial cytidine deaminase toxin enables CRISPR-free mitochondrial base editing. Nature 2020, 583, 631–637. [Google Scholar] [CrossRef]
  162. Gammage, P.A.; Moraes, C.T.; Minczuk, M. Mitochondrial Genome Engineering: The Revolution May Not Be CRISPR-Ized. Trends Genet. 2018, 34, 101–110. [Google Scholar] [CrossRef] [Green Version]
  163. Cho, S.I.; Lee, S.; Mok, Y.G.; Lim, K.; Lee, J.; Lee, J.M.; Chung, E.; Kim, J.S. Targeted A-to-G base editing in human mitochondrial DNA with programmable deaminases. Cell 2022, 185, 1764–1776 e1712. [Google Scholar] [CrossRef] [PubMed]
  164. Edwardson, J.R.; Corbett, M.K. Asexual transmission of cytoplasmic male sterility. Proc. Natl. Acad. Sci. USA 1961, 47, 390–396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Kazama, T.; Okuno, M.; Watari, Y.; Yanase, S.; Koizuka, C.; Tsuruta, Y.; Sugaya, H.; Toyoda, A.; Itoh, T.; Tsutsumi, N.; et al. Curing cytoplasmic male sterility via TALEN-mediated mitochondrial genome editing. Nat. Plants 2019, 5, 722–730. [Google Scholar] [CrossRef] [PubMed]
  166. Zsogon, A.; Cermak, T.; Naves, E.R.; Notini, M.M.; Edel, K.H.; Weinl, S.; Freschi, L.; Voytas, D.F.; Kudla, J.; Peres, L.E.P. De novo domestication of wild tomato using genome editing. Nat. Biotechnol. 2018, 36, 1211–1216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Gasparini, K.; Moreira, J.D.R.; Peres, L.E.P.; Zsogon, A. De novo domestication of wild species to create crops with increased resilience and nutritional value. Curr. Opin. Plant Biol. 2021, 60, 102006. [Google Scholar] [CrossRef] [PubMed]
  168. Allaby, R.G.; Stevens, C.J.; Kistler, L.; Fuller, D.Q. Emerging evidence of plant domestication as a landscape-level process. Trends Ecol. Evol. 2022, 37, 268–279. [Google Scholar] [CrossRef] [PubMed]
  169. 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]
  170. Ma, X.; Zhang, X.; Liu, H.; Li, Z. Highly efficient DNA-free plant genome editing using virally delivered CRISPR-Cas9. Nat. Plants 2020, 6, 773–779. [Google Scholar] [CrossRef] [PubMed]
  171. Karvelis, T.; Gasiunas, G.; Siksnys, V. Methods for decoding Cas9 protospacer adjacent motif (PAM) sequences: A brief overview. Methods 2017, 121–122, 3–8. [Google Scholar] [CrossRef] [PubMed]
  172. Chatterjee, P.; Jakimo, N.; Jacobson, J.M. Minimal PAM specificity of a highly similar SpCas9 ortholog. Sci. Adv. 2018, 4, eaau0766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Tan, J.; Zeng, D.; Zhao, Y.; Wang, Y.; Liu, T.; Li, S.; Xue, Y.; Luo, Y.; Xie, X.; Chen, L.; et al. PhieABEs: A PAM-less/free high-efficiency adenine base editor toolbox with wide target scope in plants. Plant Biotechnol. J. 2022, 20, 934–943. [Google Scholar] [CrossRef] [PubMed]
  174. Qin, R.; Li, J.; Li, H.; Zhang, Y.; Liu, X.; Miao, Y.; Zhang, X.; Wei, P. Developing a highly efficient and wildly adaptive CRISPR-SaCas9 toolset for plant genome editing. Plant Biotechnol. J. 2019, 17, 706–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Wang, M.; Xu, Z.; Gosavi, G.; Ren, B.; Cao, Y.; Kuang, Y.; Zhou, C.; Spetz, C.; Yan, F.; Zhou, X.; et al. Targeted base editing in rice with CRISPR/ScCas9 system. Plant Biotechnol. J. 2020, 18, 1645–1647. [Google Scholar] [CrossRef] [PubMed]
  176. Ren, B.; Liu, L.; Li, S.; Kuang, Y.; Wang, J.; Zhang, D.; Zhou, X.; Lin, H.; Zhou, H. Cas9-NG Greatly Expands the Targeting Scope of the Genome-Editing Toolkit by Recognizing NG and Other Atypical PAMs in Rice. Mol. Plant 2019, 12, 1015–1026. [Google Scholar] [CrossRef]
  177. Li, J.; Xu, R.; Qin, R.; Liu, X.; Kong, F.; Wei, P. Genome editing mediated by SpCas9 variants with broad non-canonical PAM compatibility in plants. Mol. Plant 2021, 14, 352–360. [Google Scholar] [CrossRef]
  178. Slaymaker, I.M.; Gao, L.; Zetsche, B.; Scott, D.A.; Yan, W.X.; Zhang, F. Rationally engineered Cas9 nucleases with improved specificity. Science 2016, 351, 84–88. [Google Scholar] [CrossRef] [Green Version]
  179. Kleinstiver, B.P.; Pattanayak, V.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Zheng, Z.; Joung, J.K. High-fidelity CRISPR-Cas9 nucleases with no detectable genome-wide off-target effects. Nature 2016, 529, 490–495. [Google Scholar] [CrossRef] [Green Version]
  180. Xu, W.; Song, W.; Yang, Y.; Wu, Y.; Lv, X.; Yuan, S.; Liu, Y.; Yang, J. Multiplex nucleotide editing by high-fidelity Cas9 variants with improved efficiency in rice. BMC Plant Biol. 2019, 19, 511. [Google Scholar] [CrossRef]
  181. Rees, H.A.; Komor, A.C.; Yeh, W.H.; Caetano-Lopes, J.; Warman, M.; Edge, A.S.B.; Liu, D.R. Improving the DNA specificity and applicability of base editing through protein engineering and protein delivery. Nat. Commun. 2017, 8, 15790. [Google Scholar] [CrossRef]
  182. Yu, Y.; Leete, T.C.; Born, D.A.; Young, L.; Barrera, L.A.; Lee, S.J.; Rees, H.A.; Ciaramella, G.; Gaudelli, N.M. Cytosine base editors with minimized unguided DNA and RNA off-target events and high on-target activity. Nat. Commun. 2020, 11, 2052. [Google Scholar] [CrossRef]
  183. Slesarenko, Y.S.; Lavrov, A.V.; Smirnikhina, S.A. Off-target effects of base editors: What we know and how we can reduce it. Curr. Genet. 2022, 68, 39–48. [Google Scholar] [CrossRef] [PubMed]
  184. 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] [PubMed]
  185. Schene, I.F.; Joore, I.P.; Oka, R.; Mokry, M.; van Vugt, A.H.M.; van Boxtel, R.; van der Doef, H.P.J.; van der Laan, L.J.W.; Verstegen, M.M.A.; van Hasselt, P.M.; et al. Prime editing for functional repair in patient-derived disease models. Nat. Commun. 2020, 11, 5352. [Google Scholar] [CrossRef] [PubMed]
  186. Anzalone, A.V.; Gao, X.D.; Podracky, C.J.; Nelson, A.T.; Koblan, L.W.; Raguram, A.; Levy, J.M.; Mercer, J.A.M.; Liu, D.R. Programmable deletion, replacement, integration and inversion of large DNA sequences with twin prime editing. Nat. Biotechnol. 2022, 40, 731–740. [Google Scholar] [CrossRef] [PubMed]
  187. Anzalone, A.V.; Koblan, L.W.; Liu, D.R. Genome editing with CRISPR-Cas nucleases, base editors, transposases and prime editors. Nat. Biotechnol. 2020, 38, 824–844. [Google Scholar] [CrossRef] [PubMed]
  188. Lorenzo, C.D.; Debray, K.; Herwegh, D.; Develtere, W.; Impens, L.; Schaumont, D.; Vandeputte, W.; Aesaert, S.; Coussens, G.; De Boe, Y.; et al. BREEDIT: A multiplex genome editing strategy to improve complex quantitative traits in maize. Plant Cell 2022, 35, 218–238. [Google Scholar] [CrossRef]
Figure 1. Mechanism of base editing work. (a) Structure and working mechanism of the cytosine base editor; (b) Structure and working mechanism of the adenine base editor; (c) Structure and editing mechanism of CGBE; (d) Window of activity for some typical CBEs, ABEs, and CGBEs.
Figure 1. Mechanism of base editing work. (a) Structure and working mechanism of the cytosine base editor; (b) Structure and working mechanism of the adenine base editor; (c) Structure and editing mechanism of CGBE; (d) Window of activity for some typical CBEs, ABEs, and CGBEs.
Cimb 45 00059 g001
Figure 2. Applications of base editing technology in crops. (a) Construction of sgRNA libraries for gene screening; (b) Single base mutations causing amino acid changes and thus developing superior traits; (c) Editing regulatory factors to regulate gene expression; (d) Introduction of stop codons in advance to regulate gene expression.
Figure 2. Applications of base editing technology in crops. (a) Construction of sgRNA libraries for gene screening; (b) Single base mutations causing amino acid changes and thus developing superior traits; (c) Editing regulatory factors to regulate gene expression; (d) Introduction of stop codons in advance to regulate gene expression.
Cimb 45 00059 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Y.; Liang, J.; Deng, B.; Jiang, Y.; Zhu, J.; Chen, L.; Li, M.; Li, J. Applications and Prospects of CRISPR/Cas9-Mediated Base Editing in Plant Breeding. Curr. Issues Mol. Biol. 2023, 45, 918-935. https://0-doi-org.brum.beds.ac.uk/10.3390/cimb45020059

AMA Style

Li Y, Liang J, Deng B, Jiang Y, Zhu J, Chen L, Li M, Li J. Applications and Prospects of CRISPR/Cas9-Mediated Base Editing in Plant Breeding. Current Issues in Molecular Biology. 2023; 45(2):918-935. https://0-doi-org.brum.beds.ac.uk/10.3390/cimb45020059

Chicago/Turabian Style

Li, Yizhen, Jing Liang, Bufang Deng, Yingli Jiang, Jingyan Zhu, Like Chen, Min Li, and Juan Li. 2023. "Applications and Prospects of CRISPR/Cas9-Mediated Base Editing in Plant Breeding" Current Issues in Molecular Biology 45, no. 2: 918-935. https://0-doi-org.brum.beds.ac.uk/10.3390/cimb45020059

Article Metrics

Back to TopTop