Next Article in Journal
Actionable Genomics in Clinical Practice: Paradigmatic Case Reports of Clinical and Therapeutic Strategies Based upon Genetic Testing
Next Article in Special Issue
Division of Labor by the HELQ, BLM, and FANCM Helicases during Homologous Recombination Repair in Drosophila melanogaster
Previous Article in Journal
Pancreatic Cancer with Mutation in BRCA1/2, MLH1, and APC Genes: Phenotype Correlation and Detection of a Novel Germline BRCA2 Mutation
Previous Article in Special Issue
The Drosophila Mutagen-Sensitivity Gene mus109 Encodes DmDNA2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

DNA Double-Strand Break Repairs and Their Application in Plant DNA Integration

1
School of Municipal and Environmental Engineering, Shandong Jianzhu University, Jinan 250101, China
2
Resources and Environment Innovation Institute, Shandong Jianzhu University, Jinan 250101, China
*
Author to whom correspondence should be addressed.
Genes 2022, 13(2), 322; https://doi.org/10.3390/genes13020322
Submission received: 19 January 2022 / Revised: 7 February 2022 / Accepted: 7 February 2022 / Published: 9 February 2022
(This article belongs to the Special Issue DNA Damage Response Mechanisms in Model Systems)

Abstract

:
Double-strand breaks (DSBs) are considered to be one of the most harmful and mutagenic forms of DNA damage. They are highly toxic if unrepaired, and can cause genome rearrangements and even cell death. Cells employ two major pathways to repair DSBs: homologous recombination (HR) and non-homologous end-joining (NHEJ). In plants, most applications of genome modification techniques depend on the development of DSB repair pathways, such as Agrobacterium-mediated transformation (AMT) and gene targeting (GT). In this paper, we review the achieved knowledge and recent advances on the DNA DSB response and its main repair pathways; discuss how these pathways affect Agrobacterium-mediated T-DNA integration and gene targeting in plants; and describe promising strategies for producing DSBs artificially, at definite sites in the genome.

1. Introduction

Double-strand breaks (DSBs) are considered to be one of the most harmful and mutagenic forms of DNA damage. They may arise as an outcome of normal cellular metabolism, but occur more frequently due to external factors such as ultraviolet (UV) radiation, ionizing radiation and genotoxic regents [1]. Being sessile organisms, plants are continually subjected to abiotic stress conditions, especially UV-B light and heavy metal pollutions, as well as to unexpected environmental changes that can also induce DNA damage [2]. DSBs are highly toxic if unrepaired, and can cause genome rearrangements and even cell death. Fortunately, cells possess highly conserved systems to recognize DSB signals and then trigger various downstream events to bring about repair. Two main DSB repair pathways are homologous recombination (HR) and non-homologous end-joining (NHEJ). The HR pathway recovers the genomic sequence precisely by using a template from the sister chromatid, the homologous chromosome or the homologous repeats in close proximity for accurate repair [3,4]. In contrast, the DSB ends are rejoined directly by the NHEJ pathway regardless of the sequence homology, leading to small deletions and insertions at the break site. HR and NHEJ are highly conserved in eukaryote cells, but their significance may be different depending on the cell type or the stage of the cell cycle. Unicellular eukaryotes mostly depend on HR to repair DSBs such as yeast Saccharomyces cerevisiae with small genomes, whereas the NHEJ pathway is the predominant one in higher eukaryotes such as humans and Arabidopsis, with large genomes containing many repeat sequences [5,6]. In plants, most applications of genome modification technique depend on the development of DSB repair pathways, such as Agrobacterium-mediated transformation (AMT) and gene targeting (GT). In this review, we summarize first the DNA DSB response and its main repair pathways. We also explore how these pathways affect Agrobacterium-mediated T-DNA integration and gene targeting in plants. In the last section, we discuss the strategies of producing site-specific DSBs artificially in the genome.

2. DNA Damage Response

The DNA damage response (DDR) is a signal transduction pathway that affects many aspects of cellular physiology (cell-cycle arrest, DNA repair, apoptosis and senescence) [7]. Key regulators of this pathway belong to the phosphatidylinositol 3-kinase-like protein kinase (PIKKs) family [8]. They control the downstream amplification of DNA damage signals by recruitment and phosphorylation of their substrates. The current understanding of the DDR mechanism in mammals is mostly dependent on the study of the three most important members of the PIKK family: ATM (Ataxia Telangiectasia Mutated), ATR (ATM and Rad-3 related) and DNA-PKcs (DNA-dependent protein kinase catalytic subunit) [9,10,11,12]. In human cells, MRE11-RAD50-NBS1 (MRN) complex and Ku70/80 heterodimer recognize DSBs signals and then trigger the activation of ATM and DNA-PKcs, respectively. Subsequently, the levels of DNA repair proteins are activated and/or induced. In addition, the appearance of areas in the genome with ssDNA as a consequence of DNA damage or repair leads to the recruitment and activation of the other master regulator, ATR. Afterwards ATM and ATR phosphorylate lots of downstream substrates, including CHK1 and CHK2, which generate a downstream amplification by protein activation and repression. These signals are subsequently transmitted to the tumor-suppressor protein p53, resulting in cell-cycle arrest and DNA repair. Another example includes the histone protein H2AX. Local phosphorylation of histone H2AX at damage sites leads to local accumulation of repair proteins, which is enhanced by ubiquitination and poly (ADP)-ribosylation of specific damage response pathway components. Multiple studies in human cells have concluded that ATM phosphorylates histone H2AX and chromatin remodeling factor KAP1 with DNA-PKcs, due to the functional redundancy between ATM and DNA-PKcs in the process [13,14,15]. However, a recent report discovered that DNA-PKcs responds to IR-mediated DSBs very quickly, and its enzymatic activity is able to initiate the DDR directly [16].
Plant DDRs are not fully understood in detail, even though the main actors of the DDR signaling pathway have been identified. So far, only a few homologous genes in plants have been characterized that are related to DDR. The Arabidopsis ATM and ATR proteins are the key DNA damage sensors for the response to DNA damage [17,18]. The MRN complex recognizes DSBs in plant cells and then activates ATM activity. Despite the conservation nature of ATM and ATR kinases in plants and animals, they seem to behave distinctively differently under DNA stress. Many genes required for the cell cycle and DNA damage checkpoints in animals have no ortholog in plants, such as CHK1, CHK2 and p53 [19]. Instead, a downstream factor called SOG1 (suppressor of γ response 1) functions as a central hub in the Arabidopsis DDR process [20,21]. As a plant-specific transcription factor, SOG1 controls the expression of most of the genes related to γ irradiation which ultimately induce cell-cycle arrest, DNA repair or programmed cell death [22,23]. In addition to being a target of SOG1, WEE1 is another critical downstream target of the ATR-ATM signaling cascades, and regulates cell-cycle arrest directly by phosphorylating and inhibiting CDKs, and/or indirectly by phosphorylating FBL17 in response to DNA damage [24,25]. Furthermore, recent findings revealed that the E2FA-RBR1 (retinoblastoma-related 1) complexes are responsible for the activation of the cell cycle checkpoint and likely function in the plant DDR in a SOG1-independent manner [26,27,28]. The role of E2FA-RBR1 complexes in DDR relies on ATM/ATR and CYCB1/CDKB activity, but the exact molecular mechanisms remain to be determined (Figure 1).

3. DSB Repair via Homologous Recombination

Homologous recombination (HR) promotes genome stability by facilitating the error-free repair of DSBs, interstrand crosslinks (ICLs), and DNA gaps, during and after DNA replication [29]. HR is a key repair pathway in the S and G2 phases of the cell cycle. HR pathway requires homologous sequences that act as a template, and the homologous sister chromatid is the preferred template in somatic cells. The information from the homologous sister chromatid is copied into, and replaces, the damaged region, resulting in precise repair. The homology search and DNA strand invasion are major steps in this process. Both are catalyzed by DNA-dependent ATPase Rad51, which can bind cooperatively to ssDNA to form helical nucleoprotein filaments [30]. HR consists of three stages, and Rad51 functions in all of them: pre-synapsis, synapsis and post-synapsis [31] (Figure 2).
At the beginning of pre-synapsis, the DSB is processed to generate ssDNA overhangs by 5′ to 3′ DNA end resection involving the Exo1 and MRN complex. In yeast, the MRX (MRE11-RAD50-XRS2) complex and Sae2 assist the short resection (50–200 nt), while more extensive resection involves either Exo1 or Sgs1 in combination with Dna2 [32]. After end resection, ssDNA overhangs are coated by RPA to protect against DNA degradation and the formation of secondary structures, which is required for the formation of competent Rad51 filaments [33]. However, RPA binding also prevents Rad51 filament assembly [34]. This inhibitory effect can be conquered by at least three different kinds of mediator proteins: Rad51 paralogs, Rad52, and BRCA2 [29]. In Arabidopsis, most of the major players in HR are identified and characterized [35,36] (Table 1).
The Rad51 filament facilitates a fast and efficient homology search and DNA strand invasion, resulting in a D-loop structure. Rad54 is required for searching for homology, stimulates DNA strand invasion by the Rad51 filament, and also functions after synapsis [37,38]. DNA synthesis is primed by the 3′ invading strand, using the donor strand as a template.
After extension of the 3′ invading strand, repair is finalized by one of at least three different sub-pathways of HR. In the classical double-strand break repair (DSBR) sub-pathway, Rad52 and Rad59 stimulate the arrest of the second end by the D-loop, whereafter two Holliday junctions (dHJ) are formed [39]. These dHJs are either dissolved into non-crossover products by a RecQ helicase, such as yeast Sgs1, or human BLM helicase, or decomposed into crossover/non-crossover products by a structure-specific endonuclease. In general, non-crossovers are much more predominant in mitotic HR [40]. Crossovers are able to produce genomic rearrangements and large-scale loss of heterozygosity (LOH) [41]. The Mph1 helicase suppresses the DSBR pathway to avoid crossover events in mitotic cells by dissociating D-loop formations, to promote annealing of the extended 3′ end to complementary sequences at the other side of the DSB, resulting in the synthesis-dependent strand annealing (SDSA) sub-pathway [42]. It seems that the Rad51 protein has some inhibitory activity that counters the capture of the second end and dHJ formation, indicating an inherent mechanistic bias toward SDSA [43]. The D-loop can convert into a replication fork in the absence of a second end, leading to the break-induced replication (BIR) sub-pathway. During BIR, the replication fork restores the integrity of a broken chromosome by copying the whole distal arm of the template chromosome, producing LOH [44]. However, no experimental evidence has been shown of the occurrence of BIR in plants [45].

4. DSB Repair via Non-Homologous End-Joining

Non-homologous end-joining (NHEJ) plays a major role in the repair of plant and mammal DSBs [5,6]. The NHEJ process seems to be relatively simple and straightforward: it rejoins broken ends directly, without the requirement of long runs of end-resection and searching for a homologous repair template. NHEJ pathways are subdivided into the canonical NHEJ (c-NHEJ) and a group of less well elucidated alternative NHEJ (a-NHEJ) pathways, which are also called microhomology-mediated end-joining (MMEJ) or polymerase theta-mediated end-joining (TMEJ) [46,47,48]. Still, many factors are required for NHEJ pathways that demand precise cooperation and timely regulation (Table 2). In general, DSB repair by NHEJ can be precise, but may also cause small nucleotide deletions and insertions at the junction, which changes the nucleotide sequence information surrounding the repair region [49]. As a result, NHEJ is considered as an error-prone DNA repair pathway. Our previous studies in Arabidopsis NHEJ-deficient mutants indicated that a-NHEJ is a more error-prone mechanism compared to c-NHEJ [50].
During the c-NHEJ pathway, DSB is recognized and bound by a Ku70/80 heterodimer to initiate the NHEJ repair [51]. The crystal structure of the human Ku heterodimer shows that the Ku heterodimer forms a ring structure on DNA broken ends in a sequence-independent manner [52]. Ku is an abundant protein that has an extraordinary affinity for dsDNA ends allowing it to quickly localize to DSBs [51,53]. The binding of Ku70/80 could protect the DSB ends from end resection, followed by the recruitment of other factors to perform end processing. Therefore, more NHEJ factors are subsequently recruited to perform end processing, including DNA-PKcs, Artemis, polynucleotide kinase/phosphatase (PNKP), the gap-filling DNA polymerases mu (Pol µ) and lambda (Pol λ), and the Mre11/Rad50/Nbs1 (MRN) complex. Arabidopsis Pol λ promotes DNA end processing in association with the XRCC4 and Ligase 4 [54]. The Ligase 4-XRCC4 complex executes final ligation of the broken ends [55].
In the absence of the Ku complex, human and plant cells still can accomplish end-joining by alternative pathways (Figure 2). The a-NHEJ pathways require elements of HR end-resection machinery, and is often associated with short tracts of microhomology different from c-NHEJ. The molecular mechanism of a-NHEJ initiation remains unclear, although both the PARP1 and the MRN complex appear to play important roles [56]. PARP1 has been well described as an important player of the BER/SSBR pathway responsible for the recruitment of the XRCC1-Ligase3 complex to stimulate repair [57]. PARP1 has been reported to compete for free DNA ends with Ku and to interact with ATM [58,59]. Therefore, PARP1 may contribute to the early damage response and is supposed to serve as a platform at the broken end, to recruit other factors. Our early results confirmed that the Arabidopsis homologs AtPARP1 and AtPARP2 are also involved in MMEJ [60]. Ku inhibits end resection, and in the absence of Ku, the MRN complex and the MRN-interacting C-terminal-binding interacting protein (CtIP) probably work together to mediate DSB resection in a-NHEJ [61]. The knockdown of Mre11 by siRNA decreased the frequency of a-NHEJ significantly without affecting the efficiency of c-NHEJ, suggesting that a-NHEJ repair of DNA DSBs requires Mre11 specifically [62]. The knockout of CtIP also results in a significant reduction in a-NHEJ [63,64]. Human Pol λ and Pol β assist MMEJ using terminal microhomology regions [65], and the interaction between PNKP and XRCC1 [66] suggests that they may engage in end processing during a-NHEJ. Furthermore, Polymerase θ (encoded by the Polq gene) which belongs to the DNA polymerase A family, plays an important role in the end processing process and is the key factor in TMEJ [48]. In studies on mus308, the Drosophila melanogaster homolog of Polq, McVey and his co-workers were the first to identify Pol θ as a factor in the a-NHEJ pathways of DSB repair [67]. Subsequently, the role of Pol θ in a-NHEJ/MMEJ has been characterized in several organisms. Biochemical studies have shown that the polymerase domain of Pol θ is able to independently carry out all of the major stages of MMEJ in vitro [68]. a-NHEJ is facilitated by sequence microhomology, which is an important signature of Pol θ-mediated end-joining. Pol θ can use a minimum of 2 bp and, optimally, 4 bp microhomology for efficient and processive DNA synthesis [69]. A second feature of TMEJ is the production of templated DNA insertions at the DSB repair junction, which can arise from sequences directly adjacent to the resected ends or other chromosomes. In addition, it has been shown that Pol θ is able to favor the a-NHEJ, by employing its ATPase activity to counteract RPA binding and promote the annealing of resected DNA substrates [70]. Recently, TMEJ has moved to the forefront of a-NHEJ, and has been raised as an indispensable player in controlling genome stability [71]. However, whether TMEJ now takes the place of MMEJ or if TMEJ is a third pathway in addition to MMEJ remains to be further determined, although several reports show that TMEJ also functions in the presence of other DSB repair pathways [72,73,74,75]. In plants, the Arabidopsis thaliana ortholog TEBICHI (Teb) is involved in DNA replication, recombination and gene expression [76]. A recent report demonstrated that Arabidopsis Pol θ participates in the repair of replication-associated DNA damage [77]. In mammalian cells, the XRCC1-Ligase3 complex seems to contribute to DSB ligation in the a-NHEJ pathway [78,79]. Since plants are lacking a Ligase 3 homolog, there must be other factors to take over ligation during a-NHEJ in plants. One of the candidates is SSB repair factor Ligase 1, which is associated in the a-NHEJ repair pathway in Arabidopsis [80]. Ligase 1 displays functional redundancy with Ligase 3 and might cooperate in a-NHEJ in mammals [81,82].

5. DSB Repair Pathway and Agrobacterium-Mediated T-DNA Integration

Agrobacterium tumefaciens is nowadays employed as a vector to create genetically modified plants. During the process of Agrobacterium-mediated genetic transformation, T-DNA is transferred from its tumor-inducing plasmid to the host cell’s nuclear genome. T-DNA is at random positions in the plant genome, which may lead to mutation and position effects altering the expression of the transgenes. Therefore, there is great interest in developing methods for the controlled and targeted integration of T-DNA. In yeast, this can be accomplished by providing a segment of yeast-homologous DNA in the T-DNA. The HR machinery of yeast then mediates integration at the homologous site [83]. In plants, homologous recombination can occur between a chromosomal locus and a homologous T-DNA introduced via Agrobacterium [84], but only with a very low efficiency. Two possible models have been recommended for T-DNA integration [85,86]. In the strand-invasion model, T-DNA integration counts on the microhomology between T-DNA and plant DNA sequences. Single-stranded T-DNA is expected to facilitate the integration [87,88,89]. In the DNA DSB repair model, the single-stranded T-DNA is first converted to a double-stranded T-DNA, whereafter this double strand form integrates into the genome at DSB sites [90]. This was supported by the fact that DSBs are preferential targets for T-DNA integration and that T-DNA can be cut by a restriction enzyme before integration [91,92]. Furthermore, by using CRISPR technology to induce DSBs, a recent finding demonstrated that T-DNA was inserted into the break sites of CRISPR/Cas9 targets with high frequency [93]. In addition, given that the microbial pathogens are capable of triggering host DNA double-strand breaks [94], the inoculation by Agrobacterium perhaps induces DSBs in host plant genomes as well, to favor the integration of T-DNA.
Although T-DNA integration has been studied for decades, its molecular mechanism has remained unclear. It should be certain that host proteins, rather than Agrobacterium proteins, are responsible for the T-DNA integration in plant genomes, which refer to the plant DNA-repair pathways. Therefore, the NHEJ pathways which have been described above are proposed to play an essential role in the integration of T-DNA to the plant genome. Earlier reports indicated that yeast (S. cerevisiae) T-DNA integration depends on NHEJ proteins, such as Ku70 and DNA ligase 4 [95,96]. NHEJ mutants in Arabidopsis were later investigated for T-DNA integration in plants. However, different research groups had conflicting results on plant T-DNA integration and revealed either no or limited negative effects [96,97,98,99]. When the multiple DNA-repair pathways were disabled at the same time, T-DNA integration was severely compromised but remained possible [100,101,102]. Nevertheless, disruption of these known repair pathways did not eliminate the end-joining, suggesting that plant T-DNA integration was mediated by other unknown proteins and pathways. Indeed, a recent study from van Kregten et al. presented that the Arabidopsis Pol θ is crucial for T-DNA integration [74]. Arabidopsis Pol θ mutants are completely recalcitrant to Agrobacterium-mediated transformation. As mentioned above, Pol θ is recognized as an important factor in the a-NHEJ of DSB repair due to its special properties of microhomology usage and template switching. These characteristics are also frequent at the T-DNA integration junctions [74,103]. ‘Filler DNA’ sequences, or alleged templated insertions, are often found at junctions and genomic sequences, which are often templated from the flank. Pol θ is proposed to extend paired 3′ overhangs at DNA synapses and use the opposing overhang as a template in trans to stabilize the DNA synapse [68]. Thus, the T-DNA left border (LB), which is a 3′ end, is the preferred substrate for Pol θ to be minimally base-paired with a 3′ end of DSB in the plant genome. Taken together, TMEJ mediates the capture of T-DNA into a DSB, providing an answer for how the T-DNA left border (LB) attaches to the plant genome (Figure 3).
Even though TMEJ explains the connecting of the T-DNA 3′ end of the integration progress, questions remain about the attachment of the T-DNA 5′ right border (RB). More recently, Nishizawa-Yokoi et al. showed that mutation of Pol θ in rice allowed stable transformation with low frequency, in which the junction fragments displayed similar characteristics to those of the wild-type plants, arguing for genes other than Pol θ in rice may also be responsible for T-DNA integration [104]. Notably, our recent results revealed that plant T-DNA integration requires Mre11 or TDP2 to remove the end protection from VirD2, to allow the capture of the T-DNA 5′ end in Arabidopsis [105]; this may fill a major gap in our understanding of T-DNA integration in plants.

6. DSB Repair Pathway and Gene Targeting

Gene targeting (GT) is a powerful genetic technique to change or replace endogenous genes depending on homologous recombination, which has been widely used to study gene function. Several approaches were established to select and detect GT events, including gene-specific selection (GSS) and positive–negative selection (PNS). In GSS schemes, an endogenous target gene is replaced by a copy of the same gene with a selectable mutation. In PNS schemes, the selection of homologous recombination relies on positive and negative selectable markers installed within and outside the homologous sequence, respectively. The PNS-based approach has proven to be very successful and is also used in plant species such as rice and Arabidopsis. GT can be achieved efficiently, especially in yeast and a few other organisms. However, GT is usually inefficient in the cells of multicellular eukaryotes, especially in those of plants, due to a much lower efficiency of HR than NHEJ [106]. The observed GT frequencies from multiple studies in plants are low and often in the 10−2 to 10−3 range [107,108,109,110,111]. In order to establish a feasible tool for GT, two options were tested to enhance the GT frequency based on the mechanism of HR. The first option was to promote the HR pathway by either increasing the synthesis of proteins involved in HR [112,113] or by inhibiting the synthesis of proteins involved in the NHEJ pathway [100]. The prevention of fungi NHEJ by deletion of Ku or Lig4 resulted in very efficient GT [96,114,115]. In plants, several reports show that GT efficiency can also be increased by blocking the NHEJ pathway or enhancing the HR pathway [116,117,118]. However, recent studies in mammalian and Arabidopsis indicate that the deficiency of Pol θ does not increase GT events [119,120]. Another approach was to enhance GT by inducing genomic DSBs at the target site, which became possible by the development of different classes of artificial nucleases (see below). In this way, the frequency of GT was increased significantly in different organisms, including plants [121,122,123,124]. GT was also achieved in Arabidopsis thaliana by expression of a site-specific endonuclease that cuts not only within the target, but also the chromosomal transgenic donor (in planta GT), leading to an excised targeting vector in each plant cell [125]. Due to recent developments, especially the application of a CRISPR/Cas nuclease system, several groups have achieved efficiency improvements of GT in plants. For instance, the group of Puchta used an egg-cell-specific promoter of the SaCas9 nuclease, which enabled them to sufficiently enhance GT efficiencies up to 1–6% in Arabidopsis [126]. Voytas and his co-workers employed geminivirus-based DNA replicons combined with Cas9, resulting in GT frequencies of ~1% in wheat [127]. The group of Levy also found that GT efficiency was strongly increased by using geminiviral replicons and a Cas9 system in tomato, with 25% in the T0 plant [128]. In addition, other works showed high GT rates of 9.1% in Arabidopsis and 8% in rice by using Cas9 and Cpf1, respectively, although they were based on small numbers of GT events [129,130]. Furthermore, several novel approaches seem to be applicable in plants to improve GT efficiency. The enhancing homology-directed repair (HDR) efficiency was observed by fusing Cas9 with a donor DNA sequence, which ultimately brought the donor DNA into close proximity to the DSB sites [131,132,133]. Moreover, it was interesting to find that the Cas9-CtIP fusion efficiently stimulated HDR after Cas9-mediated DNA cleavage [134]. Taken together, these findings might open a promising avenue for a higher efficiency of GT with the CRISPR/Cas9 system.

7. Strategies for DSB Induction

As DNA recombination events, including transgene integration and gene targeting, are increased at break sites in the genome, it has been a strategy to induce local DNA breaks to stimulate these events. Ionizing radiation (X-ray) and genotoxic chemicals (Bleomycin, MMS, etc.) were initially used to induce such DNA breaks, but as they affect the genome in an uncontrolled manner and cause mutation, this was not very successful.
Site-specific nucleases have been developed by which DSBs can be induced at a preferred site in the genome. The advent of meganucleases, such as SceI, for the first time offered the possibility to induce a DSB at a specific site(s) in a large genome. Such a local break inspired a significant increase in DNA integration [90] and in gene targeting [135]. Since then, three classes of nucleases have been used extensively: zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs) and the CRISPR/Cas (for ‘clustered regularly interspersed short palindromic repeats/CRISPR associated’) system.
ZFNs consist of zinc finger arrays fused to the nuclease domain of the type II restriction enzyme FokI. Each zinc finger typically recognizes three nucleotides, and engineered fingers have been combined to recognize specific longer DNA sequences. ZFNs function as a dimer to produce a DSB within the spacer between the binding sites of two ZFN monomers. ZFN-mediated gene modification has been reported in different eukaryotic organisms [121] and also in Arabidopsis [109,116,123]. Like ZFNs, TALENs are composed of DNA-binding domains and a FokI nuclease domain. Each binding domain includes a variable number of amino acid repeats, which are able to specifically recognize a single base pair of DNA [136]. TALENs and ZFNs make DSBs with 5′ overhangs. TALENs are considered to be more efficient, specific and reproducible, because TALENs are less affected by the context of targeting sequences than ZFNs, as shown in yeast [137], human [138] and Arabidopsis [139].
In 2012, an RNA-guided CRISPR/Cas nuclease system was described for inducing DNA DSBs at specific genomic loci [140]. CRISPR/Cas originates from a microbial adaptive immune system that uses RNA-guided nucleases to cleave the foreign invading sequences. The CRISPR/Cas9 used for DSB induction in eukaryote organisms is based on bacterial type II CRISPR/Cas systems, consisting of CRISPR-associated protein Cas9 and a single guide RNA chimera (sgRNA), which was engineered from the tracrRNA and crRNA [140]. Guided by the sgRNA via base-pairing to the target DNA sequence, both strands of the target DNA are cleaved by two endonuclease domains (HNH- and RuvC-like domains) of the Cas9 protein. The cleavage locations are also determined by a protospacer-adjacent motif (PAM) which is juxtaposed to the complementary region in the target DNA [140]. Furthermore, the CRISPR/Cas12a, formally known as Cpf1 belonging to the Class 2 type V CRISPR/Cas system, has emerged as an alternative and promising gene-editing tool with an efficiency that is at least comparable to the CRISPR/Cas9 [141,142]. Cas12a creates staggered DNA double-stranded break ends, while Cas9 produces blunt ends. The CRISPR/Cas system is markedly easier to design by changing the guide RNA sequence, compared to ZFNs and TALENs. It is highly specific and efficient for a vast number of cell types and organisms. Thus, CRISPR has quickly become a standard technique in genome engineering since its discovery. During the past ten years, the CRISPR technique has been applied successfully in various organisms and has shown incredibly fast development.

8. Concluding Remarks and Future Perspectives

Although an increasing number of proteins engaged in DNA-repair pathways have been analyzed, and their interactions investigated, in the past few decades, the repair mechanisms are not fully understood at present and require more in-depth studies. The mechanisms evolved to repair DSBs are highly conserved between organisms; however, studies in plants lag behind. These repair pathways contribute to T-DNA integration and targeted DNA insertion. The understanding of the DSB repair mechanism is beneficial to develop precise genome modification approaches, which are extremely valuable for crop plants. Furthermore, the CRISPR/Cas-mediated genome-editing tool was employed in various plants successfully in the last decade. This application not only improves our ability to deal with specific issues in fundamental research, but also adapts to creating germplasms of crop species with desired traits, and to enhancing global food security and sustainable agriculture. As we gain a deeper understanding of repair mechanisms and the improvement of nuclease-based technology, we can expect even more diversification and high-efficiency genome editing tools in the near future.

Author Contributions

H.S. and Z.L. contributed to writing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Natural Science Foundation of Shandong Province, grant number ZR2020MC058, The Introduction and Cultivation Plan for Young Innovative Talents of Colleges and Universities by the Education Department of Shandong Province, and the Youth Innovation Technology Project of Higher School in Shandong Province, grant number 2020KJF013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Sylvia de Pater and Paul JJ Hooykaas for supervision on an earlier version of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khanna, K.K.; Jackson, S.P. DNA double-strand breaks: Signaling, repair and the cancer connection. Nat. Genet. 2001, 27, 247–254. [Google Scholar] [CrossRef]
  2. Manova, V.; Gruszka, D. DNA damage and repair in plants—From models to crops. Front. Plant Sci. 2015, 6, 1–26. [Google Scholar] [CrossRef] [Green Version]
  3. Hayut, S.F.; Bessudo, C.M.; Levy, A.A. Targeted recombination between homologous chromosomes for precise breeding in tomato. Nat. Commun. 2017, 8, 1–9. [Google Scholar] [CrossRef]
  4. Siebert, R.; Puchta, H. Efficient repair of genomic double-strand breaks by homologous recombination between directly repeated sequences in the plant genome. Plant Cell 2002, 14, 1121–1131. [Google Scholar] [CrossRef] [Green Version]
  5. Stinson, B.M.; Loparo, J.J. Repair of DNA Double-Strand Breaks by the Nonhomologous End Joining Pathway. Annu. Rev. Biochem. 2021, 90, 137–164. [Google Scholar] [CrossRef]
  6. Puchta, H. The repair of double-strand breaks in plants: Mechanisms and consequences for genome evolution. J. Exp. Bot. 2005, 56, 1–14. [Google Scholar] [CrossRef] [Green Version]
  7. Spampinato, C.P. Protecting DNA from errors and damage: An overview of DNA repair mechanisms in plants compared to mammals. Cell. Mol. Life Sci. 2017, 74, 1693–1709. [Google Scholar] [CrossRef]
  8. Block, W.D.; Yu, Y.; Lees-Miller, S.P. Phosphatidyl inositol 3-kinase-like serine/threonine protein kinases (PIKKs) are required for DNA damage-induced phosphorylation of the 32 kDa subunit of replication protein A at threonine 21. Nucleic Acids Res. 2004, 32, 997–1005. [Google Scholar] [CrossRef] [Green Version]
  9. Falck, J.; Coates, J.; Jackson, S.P. Conserved modes of recruitment of ATM, ATR and DNA-PKcs to sites of DNA damage. Nature 2005, 434, 605–611. [Google Scholar] [CrossRef]
  10. Shiloh, Y. The ATM-mediated DNA-damage response: Taking shape. Trends Biochem. Sci. 2006, 31, 402–410. [Google Scholar] [CrossRef]
  11. Matsuoka, S.; Ballif, B.A.; Smogorzewska, A.; McDonald, E.R.; Hurov, K.E.; Luo, J.; Bakalarski, C.E.; Zhao, Z.; Solimini, N.; Lerenthal, Y.; et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 2007, 316, 1160–1166. [Google Scholar] [CrossRef] [Green Version]
  12. Blackford, A.N.; Jackson, S.P. ATM, ATR, and DNA-PK: The Trinity at the Heart of the DNA Damage Response. Mol. Cell 2017, 66, 801–817. [Google Scholar] [CrossRef] [Green Version]
  13. Tomimatsu, N.; Mukherjee, B.; Burma, S. Distinct roles of ATR and DNA-PKcs in triggering DNA damage responses in ATM-deficient cells. EMBO Rep. 2009, 10, 629–635. [Google Scholar] [CrossRef] [Green Version]
  14. Meyer, B.; Voss, K.O.; Tobias, F.; Jakob, B.; Durante, M.; Taucher-Scholz, G. Clustered DNA damage induces pan-nuclear H2AX phosphorylation mediated by ATM and DNA-PK. Nucleic Acids Res. 2013, 41, 6109–6118. [Google Scholar] [CrossRef] [Green Version]
  15. White, D.E.; Negorev, D.; Peng, H.; Ivanov, A.V.; Maul, G.G.; Rauscher, F.J. KAP1, a novel substrate for PIKK family members, colocalizes with numerous damage response factors at DNA lesions. Cancer Res. 2006, 66, 11594–11599. [Google Scholar] [CrossRef] [Green Version]
  16. Lu, H.; Saha, J.; Beckmann, P.J.; Hendrickson, E.A.; Davis, A.J. DNA-PKcs promotes chromatin decondensation to facilitate initiation of the DNA damage response. Nucleic Acids Res. 2019, 47, 9467–9479. [Google Scholar] [CrossRef] [Green Version]
  17. Garcia, V.; Bruchet, H.; Camescasse, D.; Granier, F.; Bouchez, D.; Tissier, A. AtATM is essential for meiosis and the somatic response to DNA damage in plants. Plant Cell 2003, 15, 119–132. [Google Scholar] [CrossRef] [Green Version]
  18. Culligan, K.M.; Robertson, C.E.; Foreman, J.; Doerner, P.; Britt, A.B. ATR and ATM play both distinct and additive roles in response to ionizing radiation. Plant J. 2006, 48, 947–961. [Google Scholar] [CrossRef]
  19. Cools, T.; De Veylder, L. DNA stress checkpoint control and plant development. Curr. Opin. Plant Biol. 2009, 12, 23–28. [Google Scholar] [CrossRef]
  20. Hu, Z.; Cools, T.; De Veylder, L. Mechanisms Used by Plants to Cope with DNA Damage. Annu. Rev. Plant Biol. 2016, 67, 439–462. [Google Scholar] [CrossRef]
  21. Yoshiyama, K.O.; Kobayashi, J.; Ogita, N.; Ueda, M.; Kimura, S.; Maki, H.; Umeda, M. ATM-mediated phosphorylation of SOG1 is essential for the DNA damage response in Arabidopsis. EMBO Rep. 2013, 14, 817–822. [Google Scholar] [CrossRef] [Green Version]
  22. Bourbousse, C.; Vegesna, N.; Law, J.A. SOG1 activator and MYB3R repressors regulate a complex DNA damage network in Arabidopsis. Proc. Natl. Acad. Sci. USA 2018, 115, E12453–E12462. [Google Scholar] [CrossRef] [Green Version]
  23. Ogita, N.; Okushima, Y.; Tokizawa, M.; Yamamoto, Y.Y.; Tanaka, M.; Seki, M.; Makita, Y.; Matsui, M.; Okamoto-Yoshiyama, K.; Sakamoto, T.; et al. Identifying the target genes of SUPPRESSOR OF γ RESPONSE 1, a master transcription factor controlling DNA damage response in Arabidopsis. Plant J. 2018, 94, 439–453. [Google Scholar] [CrossRef] [Green Version]
  24. De Schutter, K.; Joubes, J.; Cools, T.; Verkest, A.; Corellou, F.; Babiychuk, E.; Van Der Schueren, E.; Beeckman, T.; Kushnir, S.; Inze, D.; et al. Arabidopsis WEE1 kinase controls cell cycle arrest in response to activation of the DNA integrity checkpoint. Plant Cell 2007, 19, 211–225. [Google Scholar] [CrossRef] [Green Version]
  25. Pan, T.; Qin, Q.; Nong, C.; Gao, S.; Wang, L.; Cai, B.; Zhang, M.; Wu, C.; Chen, H.; Li, T.; et al. A novel WEE1 pathway for replication stress responses. Nat. Plants 2021, 7, 209–218. [Google Scholar] [CrossRef]
  26. Horvath, B.M.; Kourova, H.; Nagy, S.; Nemeth, E.; Magyar, Z.; Papdi, C.; Ahmad, Z.; Sanchez-Perez, G.F.; Perilli, S.; Blilou, I.; et al. Arabidopsis RETINOBLASTOMA RELATED directly regulates DNA damage responses through functions beyond cell cycle control. EMBO J. 2017, 36, 1261–1278. [Google Scholar] [CrossRef]
  27. Biedermann, S.; Harashima, H.; Chen, P.; Heese, M.; Bouyer, D.; Sofroni, K.; Schnittger, A. The retinoblastoma homolog RBR 1 mediates localization of the repair protein RAD 51 to DNA lesions in Arabidopsis. EMBO J. 2017, 36, 1279–1297. [Google Scholar] [CrossRef]
  28. Nisa, M.U.; Huang, Y.; Benhamed, M.; Raynaud, C. The plant DNA damage response: Signaling pathways leading to growth inhibition and putative role in response to stress conditions. Front. Plant Sci. 2019, 10, 1–12. [Google Scholar] [CrossRef] [Green Version]
  29. Heyer, W.-D.; Ehmsen, K.T.; Liu, J. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 2010, 44, 113–139. [Google Scholar] [CrossRef] [Green Version]
  30. Serra, H.; Da Ines, O.; Degroote, F.; Gallego, M.E.; White, C.I. Roles of XRCC2, RAD51B and RAD51D in RAD51-Independent SSA Recombination. PLoS Genet. 2013, 9, e1003971. [Google Scholar] [CrossRef] [Green Version]
  31. Schuermann, D.; Molinier, J.; Fritsch, O.; Hohn, B. The dual nature of homologous recombination in plants. Trends Genet. 2005, 21, 172–181. [Google Scholar] [CrossRef]
  32. Mimitou, E.P.; Symington, L.S. DNA end resection—Unraveling the tail. DNA Repair 2011, 10, 344–348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Owalczykowski, S.T.C.K. DNA annealing by Rad52 Protein is stimulated by specific interaction with the complex of replication protein A and single-stranded DNA. Proc. Natl. Acad. Sci. USA 1998, 95, 6049–6054. [Google Scholar]
  34. Sugiyama, T.; Zaitseva, E.M.; Kowalczykowski, S.C. A Single-stranded DNA-binding Protein Is Needed for Efficient Presynaptic Complex Formation by the Saccharomyces cerevisiae Rad51 Protein. J. Biol. Chem. 1997, 272, 7940–7945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Edlinger, B.; Schlögelhofer, P. Have a break: Determinants of meiotic DNA double strand break (DSB) formation and processing in plants. J. Exp. Bot. 2011, 62, 1545–1563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Mannuss, A.; Trapp, O.; Puchta, H. Gene regulation in response to DNA damage. Biochim. Biophys. Acta-Gene Regul. Mech. 2012, 1819, 154–165. [Google Scholar] [CrossRef]
  37. Mazin, A.V.; Alexeev, A.A.; Kowalczykowski, S.C. A novel function of Rad54 protein. Stabilization of the Rad51 nucleoprotein filament. J. Biol. Chem. 2003, 278, 14029–14036. [Google Scholar] [CrossRef] [Green Version]
  38. Wolner, B.; Peterson, C.L. ATP-dependent and ATP-independent roles for the Rad54 chromatin remodeling enzyme during recombinational repair of a DNA double strand break. J. Biol. Chem. 2005, 280, 10855–10860. [Google Scholar] [CrossRef] [Green Version]
  39. Wu, Y.; Sugiyama, T.; Kowalczykowski, S.C. DNA annealing mediated by Rad52 and Rad59 proteins. J. Biol. Chem. 2006, 281, 15441–15449. [Google Scholar] [CrossRef] [Green Version]
  40. Baudat, F.; de Massy, B. Regulating double-stranded DNA break repair towards crossover or non-crossover during mammalian meiosis. Chromosome Res. 2007, 15, 565–577. [Google Scholar] [CrossRef]
  41. Moynahan, M.E.; Jasin, M. Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat. Rev. Mol. Cell Biol. 2010, 11, 196–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Prakash, R.; Satory, D.; Dray, E.; Papusha, A.; Scheller, J.; Kramer, W.; Krejci, L.; Klein, H.; Haber, J.E.; Sung, P.; et al. Yeast Mph1 helicase dissociates Rad51-made D-loops: Implications for crossover control in mitotic recombination. Genes Dev. 2009, 23, 67–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Wu, Y.; Kantake, N.; Sugiyama, T.; Kowalczykowski, S.C. Rad51 protein controls Rad52-mediated DNA annealing. J. Biol. Chem. 2008, 283, 14883–14892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lydeard, J.R.; Jain, S.; Yamaguchi, M.; Haber, J.E. Break-induced replication and telomerase-independent telomere maintenance require Pol32. Nature 2007, 448, 820–823. [Google Scholar] [CrossRef] [PubMed]
  45. Schubert, I.; Schubert, V.; Fuchs, J. No evidence for “break-induced replication” in a higher plant—But break-induced conversion may occur. Front. Plant Sci. 2011, 2, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Lieber, M.R. The mechanism of double-strand DNA break repair by the nonhomologous DNA end joining pathway. Annu. Rev. Biochem. 2010, 79, 181–211. [Google Scholar] [CrossRef] [Green Version]
  47. McVey, M.; Lee, S.E. MMEJ repair of double-strand breaks (director’s cut): Deleted sequences and alternative endings. Trends Genet. 2008, 24, 529–538. [Google Scholar] [CrossRef] [Green Version]
  48. Ramsden, D.A.; Carvajal-Garcia, J.; Gupta, G.P. Mechanism, cellular functions and cancer roles of polymerase-theta-mediated DNA end joining. Nat. Rev. Mol. Cell Biol. 2022, 23, 125–140. [Google Scholar] [CrossRef]
  49. Van Gent, D.C.; van der Burg, M. Non-homologous end-joining, a sticky affair. Oncogene 2007, 26, 7731–7740. [Google Scholar] [CrossRef] [Green Version]
  50. Shen, H.; Strunks, G.D.; Klemann, B.J.P.M.; Hooykaas, P.J.J.; de Pater, S. CRISPR/Cas9-Induced Double-Strand Break Repair in Arabidopsis Nonhomologous End-Joining Mutants. G3 Genes Genomes Genet. 2017, 7, 193–202. [Google Scholar] [CrossRef] [Green Version]
  51. Waterworth, W.M.; Drury, G.E.; Bray, C.M.; West, C.E. Repairing breaks in the plant genome: The importance of keeping it together. New Phytol. 2011, 192, 805–822. [Google Scholar] [CrossRef] [PubMed]
  52. Walker, J.R.; Corpina, R.A.; Goldberg, J. Structure of the Ku heterodimer bound to DNA and its implications for double-strand break repair. Nature 2001, 412, 607–614. [Google Scholar] [CrossRef] [PubMed]
  53. Valuchova, S.; Fulnecek, J.; Prokop, Z.; Stolt-Bergner, P.; Janouskova, E.; Hofr, C.; Riha, K. Protection of arabidopsis blunt-ended telomeres is mediated by a physical association with the ku heterodimer. Plant Cell 2017, 29, 1533–1545. [Google Scholar] [CrossRef]
  54. Roy, S.; Choudhury, S.R.; Sengupta, D.N.; Das, K.P. Involvement of AtPolλ in the repair of high salt- and DNA cross-linking agent-induced double strand breaks in Arabidopsis. Plant Physiol. 2013, 162, 1195–1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. West, C.E.; Waterworth, W.M.; Jiang, Q.; Bray, C.M. Arabidopsis DNA ligase IV is induced by γ-irradiation and interacts with an Arabidopsis homologue of the double strand break repair protein XRCC4. Plant J. 2000, 24, 67–78. [Google Scholar] [CrossRef]
  56. Deriano, L.; Roth, D.B. Modernizing the nonhomologous end-joining repertoire: Alternative and classical NHEJ share the stage. Annu. Rev. Genet. 2013, 47, 433–455. [Google Scholar] [CrossRef] [PubMed]
  57. Caldecott, K.W. Single-strand break repair and genetic disease. Nat. Rev. Genet. 2008, 9, 619–631. [Google Scholar] [CrossRef] [PubMed]
  58. Haince, J.F.; McDonald, D.; Rodrigue, A.; Dery, U.; Masson, J.Y.; Hendzel, M.J.; Poirier, G.G. PARP1-dependent kinetics of recruitment of MRE11 and NBS1 proteins to multiple DNA damage sites. J. Biol. Chem. 2008, 283, 1197–1208. [Google Scholar] [CrossRef] [Green Version]
  59. Aguilar-quesada, R.; Muñoz-gámez, J.A.; Martín-Oliva, D.; Peralta, A.; Valenzuela, M.T.; Matínez-Romero, R.; Quiles-pérez, R.; Menissier-de Murcia, J.; de Murcia, G.; Ruiz de Almodóvar, M.; et al. Interaction between ATM and PARP-1 in response to DNA damage and sensitization of ATM deficient cells through PARP inhibition. BMC Mol. Biol. 2007, 8, 29. [Google Scholar] [CrossRef] [Green Version]
  60. Jia, Q.; Den Dulk-Ras, A.; Shen, H.; Hooykaas, P.J.J.; de Pater, S. Poly(ADP-ribose)polymerases are involved in microhomology mediated back-up non-homologous end joining in Arabidopsis thaliana. Plant Mol. Biol. 2013, 82, 339–351. [Google Scholar] [CrossRef]
  61. Uanschou, C.; Siwiec, T.; Pedrosa-Harand, A.; Kerzendorfer, C.; Sanchez-Moran, E.; Novatchkova, M.; Akimcheva, S.; Woglar, A.; Klein, F.; Schlögelhofer, P. A novel plant gene essential for meiosis is related to the human CtIP and the yeast COM1/SAE2 gene. EMBO J. 2007, 26, 5061–5070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zhuang, J.; Jiang, G.; Willers, H.; Xia, F. Exonuclease function of human Mre11 promotes deletional nonhomologous end joining. J. Biol. Chem. 2009, 284, 30565–30573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lee-Theilen, M.; Matthews, A.J.; Kelly, D.; Zheng, S.; Chaudhuri, J. CtIP promotes microhomology-mediated alternative end joining during class-switch recombination. Nat. Struct. Mol. Biol. 2011, 18, 75–79. [Google Scholar] [CrossRef]
  64. Zhang, Y.; Jasin, M. An essential role for CtIP in chromosomal translocation formation through an alternative end-joining pathway. Nat. Struct. Mol. Biol. 2011, 18, 80–84. [Google Scholar] [CrossRef] [PubMed]
  65. Crespan, E.; Czabany, T.; Maga, G.; Hübscher, U. Microhomology-mediated DNA strand annealing and elongation by human DNA polymerases λ and β on normal and repetitive DNA sequences. Nucleic Acids Res. 2012, 40, 5577–5590. [Google Scholar] [CrossRef] [Green Version]
  66. Weinfeld, M.; Mani, R.S.; Abdou, I.; Aceytuno, R.D.; Glover, J.N.M. Tidying up loose ends: The role of polynucleotide kinase/phosphatase in DNA strand break repair. Trends Biochem. Sci. 2011, 36, 262–271. [Google Scholar] [CrossRef] [Green Version]
  67. Chan, S.H.; Yu, A.M.; McVey, M. Dual roles for DNA polymerase theta in alternative end-joining repair of double-strand breaks in Drosophila. PLoS Genet. 2010, 6, e1001005. [Google Scholar] [CrossRef] [Green Version]
  68. Kent, T.; Chandramouly, G.; McDevitt, S.M.; Ozdemir, A.Y.; Pomerantz, R.T. Mechanism of microhomology-mediated end-joining promoted by human DNA polymerase theta. Nat. Struct. Mol. Biol. 2015, 22, 230–237. [Google Scholar] [CrossRef] [Green Version]
  69. He, P.; Yang, W. Template and primer requirements for DNA Pol θ-mediated end joining. Proc. Natl. Acad. Sci. USA 2018, 115, 7747–7752. [Google Scholar] [CrossRef] [Green Version]
  70. Mateos-Gomez, P.A.; Kent, T.; Deng, S.K.; Mcdevitt, S.; Kashkina, E.; Hoang, T.M.; Pomerantz, R.T.; Sfeir, A. The helicase domain of Polθ counteracts RPA to promote alt-NHEJ. Nat. Struct. Mol. Biol. 2017, 24, 1116–1123. [Google Scholar] [CrossRef]
  71. Schrempf, A.; Slyskova, J.; Loizou, J.I. Targeting the DNA Repair Enzyme Polymerase θ in Cancer Therapy. Trends Cancer 2021, 7, 98–111. [Google Scholar] [CrossRef] [PubMed]
  72. Koole, W.; van Schendel, R.; Karambelas, A.E.; van Heteren, J.T.; Okihara, K.L.; Tijsterman, M. A Polymerase Theta-dependent repair pathway suppresses extensive genomic instability at endogenous G4 DNA sites. Nat. Commun. 2014, 5, 3216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Mateos-Gomez, P.A.; Gong, F.; Nair, N.; Miller, K.M.; Lazzerini-Denchi, E.; Sfeir, A. Mammalian polymerase θ promotes alternative NHEJ and suppresses recombination. Nature 2015, 518, 254–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Van Kregten, M.; De Pater, S.; Romeijn, R.; Van Schendel, R.; Hooykaas, P.J.J.; Tijsterman, M. T-DNA integration in plants results from polymerase-θ-mediated DNA repair. Nat. Plants 2016, 2, 16164. [Google Scholar] [CrossRef]
  75. Wyatt, D.W.; Feng, W.; Conlin, M.P.; Yousefzadeh, M.J.; Roberts, S.A.; Mieczkowski, P.; Wood, R.D.; Gupta, G.P.; Ramsden, D.A. Essential Roles for Polymerase θ-Mediated End Joining in the Repair of Chromosome Breaks. Mol. Cell 2016, 63, 1–12. [Google Scholar] [CrossRef] [Green Version]
  76. Inagaki, S.; Nakamura, K.; Morikami, A. A link among DNA replication, recombination, and gene expression revealed by genetic and genomic analysis of TEBICHI gene of Arabidopsis thaliana. PLoS Genet. 2009, 5, e1000613. [Google Scholar] [CrossRef] [Green Version]
  77. Nisa, M.; Bergis, C.; Pedroza-Garcia, J.A.; Drouin-Wahbi, J.; Mazubert, C.; Bergounioux, C.; Benhamed, M.; Raynaud, C. The plant DNA polymerase theta is essential for the repair of replication-associated DNA damage. Plant J. 2021, 106, 1197–1207. [Google Scholar] [CrossRef]
  78. Audebert, M.; Salles, B.; Calsou, P. Involvement of poly(ADP-ribose) polymerase-1 and XRCC1/DNA ligase III in an alternative route for DNA double-strand breaks rejoining. J. Biol. Chem. 2004, 279, 55117–55126. [Google Scholar] [CrossRef] [Green Version]
  79. Simsek, D.; Brunet, E.; Wong, S.Y.W.; Katyal, S.; Gao, Y.; McKinnon, P.J.; Lou, J.; Zhang, L.; Li, J.; Rebar, E.J.; et al. DNA ligase III promotes alternative nonhomologous end-joining during chromosomal translocation formation. PLoS Genet. 2011, 7, e1002080. [Google Scholar] [CrossRef] [Green Version]
  80. Waterworth, W.M.; Kozak, J.; Provost, C.M.; Bray, C.M.; Angelis, K.J.; West, C.E. DNA ligase 1 deficient plants display severe growth defects and delayed repair of both DNA single and double strand breaks. BMC Plant Biol. 2009, 9, 79. [Google Scholar] [CrossRef] [Green Version]
  81. Arakawa, H.; Bednar, T.; Wang, M.; Paul, K.; Mladenov, E.; Bencsik-Theilen, A.A.; Iliakis, G. Functional redundancy between DNA ligases I and III in DNA replication in vertebrate cells. Nucleic Acids Res. 2012, 40, 2599–2610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Paul, K.; Wang, M.; Mladenov, E.; Bencsik-Theilen, A.; Bednar, T.; Wu, W.; Arakawa, H.; Iliakis, G. DNA ligases I and III cooperate in alternative non-homologous end-joining in vertebrates. PLoS ONE 2013, 8, e59505. [Google Scholar] [CrossRef] [PubMed]
  83. Bundock, P.; den Dulk-Ras, A.; Beijersbergen, A.; Hooykaas, P.J. Trans-kingdom T-DNA transfer from Agrobacterium tumefaciens to Saccharomyces cerevisiae. EMBO J. 1995, 14, 3206–3214. [Google Scholar] [CrossRef] [PubMed]
  84. Offringa, R.; de Groot, M.J.; Haagsman, H.J.; Does, M.P.; van den Elzen, P.J.; Hooykaas, P.J. Extrachromosomal homologous recombination and gene targeting in plant cells after Agrobacterium mediated transformation. EMBO J. 1990, 9, 3077–3084. [Google Scholar] [CrossRef] [PubMed]
  85. Tzfira, T.; Li, J.; Lacroix, B.; Citovsky, V. Agrobacterium T-DNA integration: Molecules and models. Trends Genet. 2004, 20, 375–383. [Google Scholar] [CrossRef]
  86. Gelvin, S.B. Plant Proteins Involved in Agrobacterium -Mediated Genetic Transformation. Annu. Rev. Phytopathol. 2010, 48, 45–68. [Google Scholar] [CrossRef] [Green Version]
  87. Rodenburg, K.W.; de Groot, M.J.; Schilperoort, R.A.; Hooykaas, P.J. Single-stranded DNA used as an efficient new vehicle for transformation of plant protoplasts. Plant Mol. Biol. 1989, 13, 711–719. [Google Scholar] [CrossRef]
  88. Gheysen, G.; Villarroel, R.; Van Montagu, M. Illegitimate recombination in plants: A model for T-DNA integration. Genes Dev. 1991, 5, 287–297. [Google Scholar] [CrossRef] [Green Version]
  89. Mayerhofer, R.; Koncz-kalman, Z.; Nawrath, C.; Bakkeren, G.; Crameri, A.; Angelis, K.; Redei, G.P.; Schell, J.; Hohn, B.; Koncz, C. T-DNA integration: In plants mode of illegitimate recombination. EMBO J. 1991, 10, 697–704. [Google Scholar] [CrossRef]
  90. Salomon, S.; Puchta, H. Capture of genomic and T-DNA sequences during double-strand break repair in somatic plant cells. EMBO J. 1998, 17, 6086–6095. [Google Scholar] [CrossRef]
  91. Chilton, M.-D.M.; Que, Q. Targeted integration of T-DNA into the tobacco genome at double-stranded breaks: New insights on the mechanism of T-DNA integration. Plant Physiol. 2003, 133, 956–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Tzfira, T.; Frankman, L.R.; Vaidya, M.; Citovsky, V. Site-specific integration of Agrobacterium tumefaciens T-DNA via double-stranded intermediates. Plant Physiol. 2003, 133, 1011–1023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zhang, Q.; Xing, H.L.; Wang, Z.P.; Zhang, H.Y.; Yang, F.; Wang, X.C.; Chen, Q.J. Potential high-frequency off-target mutagenesis induced by CRISPR/Cas9 in Arabidopsis and its prevention. Plant Mol. Biol. 2018, 96, 445–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Song, J.; Bent, A.F. Microbial Pathogens Trigger Host DNA Double-Strand Breaks Whose Abundance Is Reduced by Plant Defense Responses. PLoS Pathog. 2014, 10, e1004226. [Google Scholar] [CrossRef] [PubMed]
  95. Van Attikum, H.; Bundock, P.; Hooykaas, P.J. Non-homologous end-joining proteins are required for Agrobacterium T-DNA integration. EMBO J. 2001, 20, 6550–6558. [Google Scholar] [CrossRef] [Green Version]
  96. Van Attikum, H.; Hooykaas, P.J.J. Genetic requirements for the targeted integration of Agrobacterium T-DNA in Saccharomyces cerevisiae. Nucleic Acids Res. 2003, 31, 826–832. [Google Scholar] [CrossRef] [Green Version]
  97. Friesner, J.; Britt, A.B. Ku80- and DNA ligase IV-deficient plants are sensitive to ionizing radiation and defective in T-DNA integration. Plant J. 2003, 34, 427–440. [Google Scholar] [CrossRef] [Green Version]
  98. Gallego, M.E.; Bleuyard, J.-Y.; Daoudal-Cotterell, S.; Jallut, N.; White, C.I. Ku80 plays a role in non-homologous recombination but is not required for T-DNA integration in Arabidopsis. Plant J. 2003, 35, 557–565. [Google Scholar] [CrossRef]
  99. Li, J.; Vaidya, M.; White, C.; Vainstein, A.; Citovsky, V.; Tzfira, T. Involvement of KU80 in T-DNA integration in plant cells. Proc. Natl. Acad. Sci. USA 2005, 102, 19231–19236. [Google Scholar] [CrossRef] [Green Version]
  100. Jia, Q.; Bundock, P.; Hooykaas, P.J.J.; de Pater, S. Agrobacterium tumefaciens T-DNA Integration and Gene Targeting in Arabidopsis thaliana Non-Homologous End-Joining Mutants. J. Bot. 2012, 2012, 1–13. [Google Scholar] [CrossRef] [Green Version]
  101. Mestiri, I.; Norre, F.; Gallego, M.E.; White, C.I. Multiple host-cell recombination pathways act in Agrobacterium-mediated transformation of plant cells. Plant J. 2014, 77, 511–520. [Google Scholar] [CrossRef] [PubMed]
  102. Park, S.-Y.; Vaghchhipawala, Z.; Vasudevan, B.; Lee, L.-Y.; Shen, Y.; Singer, K.; Waterworth, W.M.; Zhang, Z.J.; West, C.E.; Mysore, K.S.; et al. Agrobacterium T-DNA integration into the plant genome can occur without the activity of key non-homologous end-joining proteins. Plant J. 2015, 81, 934–946. [Google Scholar] [CrossRef] [PubMed]
  103. Kleinboelting, N.; Huep, G.; Appelhagen, I.; Viehoever, P.; Li, Y.; Weisshaar, B. The Structural Features of Thousands of T-DNA Insertion Sites Are Consistent with a Double-Strand Break Repair-Based Insertion Mechanism. Mol. Plant 2015, 8, 1651–1664. [Google Scholar] [CrossRef] [PubMed]
  104. Nishizawa-Yokoi, A.; Saika, H.; Hara, N.; Lee, L.Y.; Toki, S.; Gelvin, S.B. Agrobacterium T-DNA integration in somatic cells does not require the activity of DNA polymerase θ. New Phytol. 2021, 229, 2859–2872. [Google Scholar] [CrossRef] [PubMed]
  105. Kralemann, L.; de Pater, S.; Shen, H.; Kloet, S.; van Schendel, R.; Hooykaas, P.J.J.; Tijsterman, M. T-DNA integration in plants requires MRE11- or TDP2-mediated removal of the 5 ’ bound Agrobacterium protein VirD2. Nat. Plants 2022. submitted. [Google Scholar]
  106. Knoll, A.; Fauser, F.; Puchta, H. DNA recombination in somatic plant cells: Mechanisms and evolutionary consequences. Chromosom. Res. 2014, 22, 191–201. [Google Scholar] [CrossRef]
  107. 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]
  108. Ayar, A.; Wehrkamp-Richter, S.; Laffaire, J.B.; Le Goff, S.; Levy, J.; Chaignon, S.; Salmi, H.; Lepicard, A.; Sallaud, C.; Gallego, M.E.; et al. Gene targeting in maize by somatic ectopic recombination. Plant Biotechnol. J. 2013, 11, 305–314. [Google Scholar] [CrossRef] [Green Version]
  109. 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] [Green Version]
  110. Watanabe, K.; Breier, U.; Hensel, G.; Kumlehn, J.; Schubert, I.; Reiss, B. Stable gene replacement in barley by targeted double-strand break induction. J. Exp. Bot. 2016, 67, 1433–1445. [Google Scholar] [CrossRef] [Green Version]
  111. De Pater, S.; Klemann, B.J.P.M.; 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, 2–11. [Google Scholar] [CrossRef] [PubMed]
  112. Reiss, B.; Schubert, I.; Köpchen, K.; Wendeler, E.; Schell, J.; Puchta, H. RecA stimulates sister chromatid exchange and the fidelity of double-strand break repair, but not gene targeting, in plants transformed by Agrobacterium. Proc. Natl. Acad. Sci. USA 2000, 97, 3358–3363. [Google Scholar] [CrossRef] [PubMed]
  113. Shaked, H.; Melamed-Bessudo, C.; Levy, A.A. High-frequency gene targeting in Arabidopsis plants expressing the yeast RAD54 gene. Proc. Natl. Acad. Sci. USA 2005, 102, 12265–12269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kooistra, R.; Hooykaas, P.J.J.; Steensma, H.Y. Efficient gene targeting in Kluyveromyces lactis. Yeast 2004, 21, 781–792. [Google Scholar] [CrossRef]
  115. De Boer, P.; Bastiaans, J.; Touw, H.; Kerkman, R.; Bronkhof, J.; van den Berg, M.; Offringa, R. Highly efficient gene targeting in Penicillium chrysogenum using the bi-partite approach in Δlig4 or Δku70 mutants. Fungal Genet. Biol. 2010, 47, 839–846. [Google Scholar] [CrossRef]
  116. Qi, Y.; Zhang, Y.; Zhang, F.; Baller, J.A.; Cleland, S.C.; Ryu, Y.; Starker, C.G.; Voytas, D.F. Increasing frequencies of site-specific mutagenesis and gene targeting in Arabidopsis by manipulating DNA repair pathways. Genome Res. 2013, 23, 547–554. [Google Scholar] [CrossRef] [Green Version]
  117. Endo, M.; Mikami, M.; Toki, S. Biallelic gene targeting in rice. Plant Physiol. 2016, 170, 667–677. [Google Scholar] [CrossRef] [Green Version]
  118. Even-Faitelson, L.; Samach, A.; Melamed-Bessudo, C.; Avivi-Ragolsky, N.; Levy, A.A. Localized egg-cell expression of effector proteins for targeted modification of the Arabidopsis genome. Plant J. 2011, 68, 929–937. [Google Scholar] [CrossRef]
  119. Saito, S.; Maeda, R.; Adachi, N. Dual loss of human POLQ and LIG4 abolishes random integration. Nat. Commun. 2017, 8, 16112. [Google Scholar] [CrossRef] [Green Version]
  120. Van Tol, N.; van Schendel, R.; Bos, A.; van Kregten, M.; de Pater, S.; Hooykaas, P.J.J.; Tijsterman, M. Gene targeting in polymerase theta-deficient Arabidopsis thaliana. Plant J. 2022, 109, 112–125. [Google Scholar] [CrossRef]
  121. Urnov, F.D.; Rebar, E.J.; Holmes, M.C.; Zhang, H.S.; Gregory, P.D. Genome editing with engineered zinc finger nucleases. Nat. Rev. Genet. 2010, 11, 636–646. [Google Scholar] [CrossRef] [PubMed]
  122. 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]
  123. De Pater, S.; Neuteboom, L.W.; Pinas, J.E.; Hooykaas, P.J.J.; Van Der Zaal, B.J. ZFN-induced mutagenesis and gene-targeting in Arabidopsis through Agrobacterium-mediated floral dip transformation. Plant Biotechnol. J. 2009, 7, 821–835. [Google Scholar] [CrossRef]
  124. Wang, M.; Lu, Y.; Botella, J.R.; Mao, Y.; Hua, K.; Zhu, J. kang Gene Targeting by Homology-Directed Repair in Rice Using a Geminivirus-Based CRISPR/Cas9 System. Mol. Plant 2017, 10, 1007–1010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Fauser, F.; Roth, N.; Pacher, M.; Ilg, G.; Sanchez-Fernandez, R.; Biesgen, C.; Puchta, H. In planta gene targeting. Proc. Natl. Acad. Sci. USA 2012, 109, 7535–7540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wolter, F.; Klemm, J.; Puchta, H. Efficient in planta gene targeting in Arabidopsis using egg cell-specific expression of the Cas9 nuclease of Staphylococcus aureus. Plant J. 2018, 94, 735–746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Gil-Humanes, J.; Wang, Y.; Liang, Z.; Shan, Q.; Ozuna, C.V.; Sánchez-León, S.; Baltes, N.J.; Starker, C.; Barro, F.; Gao, C.; et al. High-efficiency gene targeting in hexaploid wheat using DNA replicons and CRISPR/Cas9. Plant J. 2017, 89, 1251–1262. [Google Scholar] [CrossRef] [Green Version]
  128. Dahan-Meir, T.; Filler-Hayut, S.; Melamed-Bessudo, C.; Bocobza, S.; Czosnek, H.; Aharoni, A.; Levy, A.A. Efficient in planta gene targeting in tomato using geminiviral replicons and the CRISPR/Cas9 system. Plant J. 2018, 95, 5–16. [Google Scholar] [CrossRef] [Green Version]
  129. Miki, D.; Zhang, W.; Zeng, W.; Feng, Z.; Zhu, J.K. CRISPR/Cas9-mediated gene targeting in Arabidopsis using sequential transformation. Nat. Commun. 2018, 9, 1–9. [Google Scholar] [CrossRef]
  130. 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, 1–6. [Google Scholar] [CrossRef] [Green Version]
  131. Ma, M.; Zhuang, F.; Hu, X.; Wang, B.; Wen, X.Z.; Ji, J.F.; Xi, J.J. Efficient generation of mice carrying homozygous double-floxp alleles using the Cas9-Avidin/Biotin-donor DNA system. Cell Res. 2017, 27, 578–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Aird, E.J.; Lovendahl, K.N.; St Martin, A.; Harris, R.S.; Gordon, W.R. Increasing Cas9-mediated homology-directed repair efficiency through covalent tethering of DNA repair template. Commun. Biol. 2018, 1, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Savic, N.; Ringnalda, F.C.A.S.; Lindsay, H.; Berk, C.; Bargsten, K.; Li, Y.; Neri, D.; Robinson, M.D.; Ciaudo, C.; Hall, J.; et al. Covalent linkage of the DNA repair template to the CRISPR-Cas9 nuclease enhances homology-directed repair. eLife 2018, 7, 1–18. [Google Scholar] [CrossRef]
  134. Charpentier, M.; Khedher, A.H.Y.; Menoret, S.; Brion, A.; Lamribet, K.; Dardillac, E.; Boix, C.; Perrouault, L.; Tesson, L.; Geny, S.; et al. CtIP fusion to Cas9 enhances transgene integration by homology-dependent repair. Nat. Commun. 2018, 9, 1–11. [Google Scholar] [CrossRef] [PubMed]
  135. Smih, F.; Rouet, P.; Romanienko, P.J.; Jasin, M. Double-strand breaks at the target locus stimulate gene targeting in embryonic stem cells. Nucleic Acids Res. 1995, 23, 5012–5019. [Google Scholar] [CrossRef] [PubMed]
  136. Joung, J.K.; Sander, J.D. TALENs: A widely applicable technology for targeted genome editing. Nat. Rev. Mol. Cell Biol. 2013, 14, 49–55. [Google Scholar] [CrossRef] [Green Version]
  137. 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, 756–761. [Google Scholar] [CrossRef] [Green Version]
  138. Zhang, F.; Cong, L.; Lodato, S.; Kosuri, S.; Church, G.M.; Arlotta, P. Efficient construction of sequence-specific TAL effectors for modulating mammalian transcription. Nat. Biotechnol. 2011, 29, 149–153. [Google Scholar] [CrossRef] [Green Version]
  139. Cermak, T.; Doyle, E.; Christian, M. Efficient design and assembly of custom TALEN and other TAL effector-based constructs for DNA targeting. Nucleic Acids Res. 2011, 39, e82. [Google Scholar] [CrossRef] [Green Version]
  140. 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]
  141. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; Van Der Oost, J.; Regev, A.; et al. Cpf1 Is a Single RNA-Guided Endonuclease of a Class 2 CRISPR-Cas System. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Paul, B.; Montoya, G. CRISPR-Cas12a: Functional overview and applications. Biomed. J. 2020, 43, 8–17. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of DNA Double-strand break (DSB) damage response in plants. MRN (MRE11-RAD50-NBS1) complex first recognizes the DNA DSB resulting in ATM (Ataxia Telangiectasia Mutated) activation. Subsequently, the ATM amplify signals by phosphorylation of downstream substrates, such as SOG1 (suppressor of γ response 1). The SOG1 functions as a central hub and controls the expression of hundreds of genes, which ultimately induce cell-cycle arrest, DNA repair or programmed cell death.
Figure 1. Overview of DNA Double-strand break (DSB) damage response in plants. MRN (MRE11-RAD50-NBS1) complex first recognizes the DNA DSB resulting in ATM (Ataxia Telangiectasia Mutated) activation. Subsequently, the ATM amplify signals by phosphorylation of downstream substrates, such as SOG1 (suppressor of γ response 1). The SOG1 functions as a central hub and controls the expression of hundreds of genes, which ultimately induce cell-cycle arrest, DNA repair or programmed cell death.
Genes 13 00322 g001
Figure 2. The major pathways of DNA DSB repair: (a) During the c-NHEJ, DSB is recognized and bound by Ku70/80. The binding of Ku70/80 can protect the DSB ends from end resection, followed by the recruitment of other factors to perform end processing. The Ligase4-XRCC4 complex executes the final ligation step; (b) in a-NHEJ (MMEJ or TMEJ) pathways, PARP1 is supposed to serve as a platform at the broken end to recruit other factors, including the DNA polymerase θ (Pol θ) which utilizes short microhomologies (indicated as pink boxes) for efficient and processive DNA synthesis. The microhomology-mediated joints between the two DNA ends are stabilized by Pol θ and work as primers for gap filling, while the XRCC1-Ligase3 (Ligase1) complex is responsible for the final ligation step; (c) HR is initiated by the long-range DNA end resection involving Exo1 and MRN complex. Subsequently, ssDNA overhangs are coated by RPA for protection against winding of the DNA. The Rad51 filament then facilitates a fast and efficient homology search and DNA strand invasion, resulting in a D-loop structure. One of three sub-pathways of HR complete the repair in the end: synthesis-dependent strand annealing (SDSA), double-strand break repair (DSBR) or break-induced replication (BIR) (does not exist in plants).
Figure 2. The major pathways of DNA DSB repair: (a) During the c-NHEJ, DSB is recognized and bound by Ku70/80. The binding of Ku70/80 can protect the DSB ends from end resection, followed by the recruitment of other factors to perform end processing. The Ligase4-XRCC4 complex executes the final ligation step; (b) in a-NHEJ (MMEJ or TMEJ) pathways, PARP1 is supposed to serve as a platform at the broken end to recruit other factors, including the DNA polymerase θ (Pol θ) which utilizes short microhomologies (indicated as pink boxes) for efficient and processive DNA synthesis. The microhomology-mediated joints between the two DNA ends are stabilized by Pol θ and work as primers for gap filling, while the XRCC1-Ligase3 (Ligase1) complex is responsible for the final ligation step; (c) HR is initiated by the long-range DNA end resection involving Exo1 and MRN complex. Subsequently, ssDNA overhangs are coated by RPA for protection against winding of the DNA. The Rad51 filament then facilitates a fast and efficient homology search and DNA strand invasion, resulting in a D-loop structure. One of three sub-pathways of HR complete the repair in the end: synthesis-dependent strand annealing (SDSA), double-strand break repair (DSBR) or break-induced replication (BIR) (does not exist in plants).
Genes 13 00322 g002
Figure 3. Simplified model of T-DNA integration: polymerase θ is required for capture of the T-DNA’s 3′ end; MRN or TDP2 helps to remove VirD2 for genomic capture of the 5′ end.
Figure 3. Simplified model of T-DNA integration: polymerase θ is required for capture of the T-DNA’s 3′ end; MRN or TDP2 helps to remove VirD2 for genomic capture of the 5′ end.
Genes 13 00322 g003
Table 1. HR proteins in human, yeast and Arabidopsis.
Table 1. HR proteins in human, yeast and Arabidopsis.
Homo sapiensSaccharomyces
cerevisiae
Arabidopsis
thaliana
Arabidopsis
Gene Number
Function
Rad51Rad51Rad51At5g20850RecA homologue
Strand invasion
MRN complex: Mre11-Rad50-Nbs1MRX complex: Mre11-Rad50-Xrs2MRN complex: Mre11-Rad50-Nbs1At5g54260
At2g31970
At3g02680
DNA binding
Nuclease activities
DSB end processing
DNA-damage checkpoints
CtIPSae2Com1At3g52115DSB end processing
DNA strand transition
Exo1Exo1Exo1A
Exo1B
At1g29630
At1g18090
DSB end processing
BLMSgs1RecQ4A
RecQ4B
At1g10930
At1g60930
DSB end processing
RecQ helicases
RPA1
RPA2
RPA3
RPA1
RPA2
RPA3
RPA1
RPA2
RPA3
At2g06510
At4g19130
At5g45400
At5g08020
At5g61000
ssDNA binding
Rad51B-Rad51C
Rad51C-XRCC3
Rad51D-XRCC2
Rad55-Rad57Rad51B-Rad51C
Rad51C-XRCC3
Rad51D-XRCC2
At2g28560
At2g45280
At5g57450
At1g07745
At5g64520
ssDNA binding
Recombination mediator
Rad52Rad52Rad52At1g71310
At5g47870
ssDNA binding and annealing
Recombination mediator
Interacts with Rad51 and RPA
BRCA11BRCA1At4g21070Checkpoint mediator
Recombination mediator
BRCA21BRCA2-1
BRCA2-2
At5g01630
At4g00020
Recombination mediator
Rad54Rad54Rad54At3g19210ATP-dependent dsDNA translocase
Stimulates the D-loop reaction
FancMMph1FancMAt1g35530Helicase activity
Dissociates D-loop formation and facilitates single-strand annealing
1 No yeast equivalent has been identified.
Table 2. NHEJ proteins in human, yeast and Arabidopsis.
Table 2. NHEJ proteins in human, yeast and Arabidopsis.
Homo sapiensSaccharomyces
cerevisiae
Arabidopsis
thaliana
Arabidopsis
Gene Number
Function
Ku70/Ku80Ku70/Ku80Ku70/Ku80At1g16970
At1g48050
DSB end binding and protection
DNA-PKcs11 protein kinase
ArtemisSnm1/PSO2Snm1At3g26680DNA end processing
MRN complex: Mre11-Rad50-Nbs1MRX complex: Mre11-Rad50-Xrs2MRN complex: Mre11-Rad50-Nbs1At5g54260
At2g31970
At3g02680
DNA binding
Nuclease activities
DSB ends processing
DNA-damage checkpoints
PNKPTpp1ZDPAt3g14890DNA end processing
Pol λ1Pol λAt1g10520DNA polymerase
DNA end processing
53BP1Rad9Rad9At3g05480DNA end processing
DNA ligase IVDnl4lig4At5g57160ATP-dependent DNA ligase
XRCC4Lif1XRCC4At3g23100complex with lig4
XLF/CernunnosNej11 complex with lig4/XRCC4
Parp11Parp1At2g31320DNA end binding
NAD + ADP-ribosyltransferase
Parp21Parp2At4g02390DNA end binding
NAD + ADP-ribosyltransferase
Parp31Parp3At5g22470DNA end binding
NAD + ADP-ribosyltransferase
CtIPSae2Com1At3g52115DNA end processing
DNA ligase III11 ATP-dependent DNA ligase
XRCC11XRCC1At1g80420complex with lig3
Pol Q1Pol θ (Tebichi)At4g32700DNA polymerase
DNA end processing
1 No yeast or Arabidopsis equivalent has been identified.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shen, H.; Li, Z. DNA Double-Strand Break Repairs and Their Application in Plant DNA Integration. Genes 2022, 13, 322. https://doi.org/10.3390/genes13020322

AMA Style

Shen H, Li Z. DNA Double-Strand Break Repairs and Their Application in Plant DNA Integration. Genes. 2022; 13(2):322. https://doi.org/10.3390/genes13020322

Chicago/Turabian Style

Shen, Hexi, and Zhao Li. 2022. "DNA Double-Strand Break Repairs and Their Application in Plant DNA Integration" Genes 13, no. 2: 322. https://doi.org/10.3390/genes13020322

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