Next Article in Journal
A Novel Cell Delivery System Exploiting Synergy between Fresh Titanium and Fibronectin
Next Article in Special Issue
The Role of ATR Inhibitors in Ovarian Cancer: Investigating Predictive Biomarkers of Response
Previous Article in Journal
Human Basal and Suprabasal Keratinocytes Are Both Able to Generate and Maintain Dermo–Epidermal Skin Substitutes in Long-Term In Vivo Experiments
Previous Article in Special Issue
Telomere Maintenance and the cGAS-STING Pathway in Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exonucleases: Degrading DNA to Deal with Genome Damage, Cell Death, Inflammation and Cancer

by
Joan Manils
1,2,†,
Laura Marruecos
3,† and
Concepció Soler
2,4,*
1
Serra Húnter Programme, Immunology Unit, Department of Pathology and Experimental Therapy, School of Medicine, Universitat de Barcelona, Feixa Llarga s/n, 08907 L’Hospitalet de Llobregat, Spain
2
Immunity, Inflammation and Cancer Group, Oncobell Program, Institut d’Investigació Biomèdica de Bellvitge—IDIBELL, 08907 L’Hospitalet de Llobregat, Spain
3
Breast Cancer Laboratory, Cancer Biology and Stem Cells Division, The Walter and Eliza Hall Institute of Medical Research, Parkville, VIC 3052, Australia
4
Immunology Unit, Department of Pathology and Experimental Therapy, School of Medicine, Universitat de Barcelona, 08007 Barcelona, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2022, 11(14), 2157; https://doi.org/10.3390/cells11142157
Submission received: 30 May 2022 / Revised: 30 June 2022 / Accepted: 7 July 2022 / Published: 9 July 2022
(This article belongs to the Special Issue DNA Damage Response Regulation and Cancer)

Abstract

:
Although DNA degradation might seem an unwanted event, it is essential in many cellular processes that are key to maintaining genomic stability and cell and organism homeostasis. The capacity to cut out nucleotides one at a time from the end of a DNA chain is present in enzymes called exonucleases. Exonuclease activity might come from enzymes with multiple other functions or specialized enzymes only dedicated to this function. Exonucleases are involved in central pathways of cell biology such as DNA replication, repair, and death, as well as tuning the immune response. Of note, malfunctioning of these enzymes is associated with immune disorders and cancer. In this review, we will dissect the impact of DNA degradation on the DNA damage response and its links with inflammation and cancer.

1. Role of Exonucleases

The description by Watson and Crick of the structure of DNA in the early 1950s [1] led to a revolution in molecular biology. The capacity of DNA to store and replicate the information required for cells and organisms to live was later discovered. Nowadays, everyone knows that DNA is the essential genetic material containing the map and instructions of who we are. DNA is so important that eukaryotic cells dedicate a whole lipidic fence (nuclear envelope) and heavy compaction to protect it. It might be difficult to understand then, why a cell would want to degrade these precious nucleic acids. However, cells have hundreds of different proteins with the capacity for cutting nucleic acids, and such an investment in this activity indicates that eliminating DNA is vital.
DNA is made of two chains of polynucleotides. The building bricks of DNA, the nucleotides, contain three components, a sugar attached to a base containing nitrogen (adenine (A), thymine (T), guanine (G), or cytosine (C)) and a phosphate group that, through phosphodiester bonds, interlinks the 5′-phosphate end of one sugar to the 3′-hydroxyl end of the next sugar, forming the polynucleotide chains. Phosphodiester (P-O) bonds are among the most versatile and stable biochemical bridges between biomolecules [2]. However, nucleases are able to cleave one of the two phosphodiester bonds that link adjacent sugars. There are multiple types with multiple functions, but grossly one can divide nucleases according to the type of substrate they cleave (RNAses [3] or DNAses [4]) and wherein the nucleic acid chain they perform the cut (endo- or exonucleases). While endonucleases cut the P-O bond from inside the polynucleotide chain generating two oligonucleotides and can be sequence- or structure-specific [5], exonucleases hydrolyze the bonds from the outer ends of the chain. Exonucleases can sequentially cleave P-O bonds from 3′-OH or from 5′-P of a single or double DNA chain in a nonspecific manner, generating individual nucleotide monophosphates [6].
The molecular event of a chemical modification of the DNA structure triggers signalling cascades that ultimately produce a cellular response. To maintain genome integrity, cells have a DNA damage response (DDR) mechanism, a multiple pathway response that integrates DNA damage sensing, DNA repair machinery, halting of the cell cycle and if repair is not possible, cell death [7]. Lesions in DNA are sensed by specialized proteins such as ATM, DNA-PK, and ATR [8], which act depending on the type of lesion and the cell cycle phase. While the different factors required to repair a specific DNA lesion are being activated and recruited to the damaged sites, p53 protein receives the signals to stop the cell cycle [9], thus preventing the transmission of DNA lesions to the daughter cell. Exonuclease activity is important in all steps in this process, from DNA sensing and repair to cell death.
Distinct exonucleases, such as APE1 [10], EXO1 [11], FAN1 [12], and FEN1 [13], are important components of several DNA repair pathways, including base excision repair (BER), nucleotide excision repair (NER), mismatch repair (MMR), non-homologous end joining (NHEJ), homologous recombination (HR), single-strand break repair (SSBR), post-replication repair (PRR) or DNA damage tolerance (DDT), interstrand cross-link repair (ICL), stalled replication fork and hairpin structure repair, as well as polymerase proofreading, as detailed below. Their ability to cleave DNA allows the elimination of damaged or mismatched nucleotides, which facilitates subsequent insertion of the correct base [14].
Some other exonucleases take part in apoptosis. Apoptosis occurs in normal development, cell turnover, and lymphocyte maturation but also in response to stress such as infection or excessive DNA damage. During apoptosis, DNA is condensed and fragmented [15] to facilitate digestion by engulfing macrophages [16]. For instance, the apoptosis enhancing nuclease (AEN), an exonuclease [17] transcribed by activated p53, is required for p53-induced apoptosis [18]. TREX1 expression increases upon genotoxic damage [19] and contributes to cell death induced by GzmA [20]. GzmA is part of the SET complex, which is released by cytotoxic cells to degrade DNA, prevent its repair and ensure death [21]. Similarly, the keratinocyte-specific TREX2 exonuclease promotes the passage of UVB-irradiated keratinocytes to late non-reversible apoptotic stages [22]. Other exonucleases participate in the degradation of DNA upon apoptosis activation, such as ARTEMIS [23], FEN1 [13] and APE1 [10].
Foreign and self-nucleic acids pose a threat to the organism, and exonucleases play an important role in tuning the innate immune response. By degrading DNA from pathogens, exonucleases control both invader infection and type I interferon (IFN) immune responses that are driven by DNA-sensing proteins [24]. Because nucleic acid sensors can also recognize endogenous DNA [25], nucleases are pivotal in removing excessive endogenous DNA to prevent detection. Exonucleases like TREX1 in the cytosol and PLD3 and PLD4 in the endolysosomes regulate cytosolic cGAS/STING activation and endosomal TLR nucleic acid-sensing, respectively [26,27], thereby limiting DNA-driven autoimmune diseases, such as rheumatoid arthritis and lupus [28,29]. Of note, autoimmunity may also be a risk factor for cancer [30,31].
Hence, exonuclease activity might come from proteins with single or multiple functional domains, such as apoptotic nucleases and DNA polymerases respectively. As stated above, nucleotide cleavage by exonucleases is important in many and quite different cell processes, from DNA synthesis/repair to DNA degradation during cell death, including DNA-driven inflammatory responses, maintaining genome stability, and ensuring the viability of the organism (Figure 1) [32,33,34].
Here, we focus on proteins with robust exonuclease activity and their role in the DDR and cancer. Thus, we comment on AEN, APE1, ARTEMIS (DCLRE1C), EXD2, EXO1, EXOG, FAN1, FEN1, MRE11A, p53, PLD3, PDL4, POLD1, POLE, RAD9A, TREX1, TREX2, and WRN, most of them included in the recently curated list of DNA Damage Repair genes in cancer [35].
To ascertain functional interactions, we performed an analysis of the 18 above-mentioned exonucleases using the STRING database of known and predicted protein–protein interactions (Figure 2). Ten exonucleases (EXO1, WRN, p53, MRE11, RAD9A, DCLRE1C, FEN1, APEX1, POLE y POLD1) were interconnected, indicating that interactions between them have been described at least in curated databases, experiments or in the literature, and functionally associated. All these exonucleases were significantly associated with the general GO process “DNA metabolic process” (dark blue) and most of them participate in DNA repair pathways.

2. Exonucleases and Cancer

2.1. AEN

Apoptosis enhancing nuclease (AEN), also known as ISG20L1, is an exonuclease that is highly efficient at processing 3′ DNA ends [17]. It is transcribed by activated p53 and promotes both single- and double-stranded DNA and RNA digestion to amplify apoptosis. If absent, cells are resistant to this type of cell death [18]. Importantly, expression of AEN not only promotes but is also required for autophagy [36].
AEN expression is upregulated in human peripheral blood mononuclear cells upon low-energy X-ray exposure during dual-energy computed tomography (DECT) [37] and cyclophosphamide treatment, stimulating the proinflammatory cell death of both tumour and blood cells and thus enhancing the efficacy of immunotherapy [38]. Moreover, bufalin also induced the expression of AEN in lung cancer cells in vitro [39]. AEN was included in a marker signature that can identify patients with a high risk of biochemical recurrence in prostate cancer (Table 1). High levels of gene expression, together with other genes, can predict recurrence [40]. Similarly, enhanced expression of AEN was used as a prognostic marker in an RNA-binding protein signature for colorectal cancer [41]. Given its response to genotoxins and its links to p53 and apoptosis, it is rather surprising that there is little existing knowledge on the role of AEN in cancer.

2.2. APE1

Apurinic/apyrimidinic endonuclease 1 (APE1), APEX1 or reduction-oxidation factor (Ref1), is a multifunctional enzyme, its main function being to incise the phosphodiester bond immediately 5′ to apurinic/apyrimidinic (AP) sites to generate single-strand breaks (SSBs). However, this protein also possesses 3′–5′ exonuclease activity [10]. It is involved in maintaining genome stability, participating in several DNA repair pathways such as tri-nucleotide repair (TNR) by the removal of hairpin structures and BER, digestion of matched and mismatched 3′ ends of duplex DNA structures, and the recognition of SSBs to induce their repair, and in apoptosis by exonucleolytic digestion of chromosomal fragments. It also prevents the formation of double-strand breaks (DSBs) during the repair of bi-stranded clustered DNA damage by nucleotide incision repair (NIR), which repairs oxidative damage in nucleotides, and interacts with POLB [10] to carry out proofreading.
Although as an endonuclease it is highly specific for AP sites, as an exonuclease it can recognize a wide range of abnormal nucleotides that are generated by oxidative stress, ionizing radiation (IR), or drug treatments [42], such as anti-cancer and anti-viral therapies. Therefore, inhibiting APE1 is an attractive approach for killing tumour cells; in fact, some APE1 inhibitors are already in clinical trials [43,44,45]. Reducing the levels of APE1 sensitizes the cells to PARP inhibitor treatment [46], hence combined therapy with PARP and APE1 inhibitors has been suggested to be highly effective in breast cancer.
Cell lines deficient for APE1 accumulate DNA damage and induce stress response pathways such as senescence [47,48,49]. Ape1 knockout mice (Table 2) showed embryonic lethality [50]. However, conditional deletion of the gene early after birth induced impaired growth, reduced organ size, and increased cellular senescence in tissues like skin or colon [49]. These mice also showed an accumulation of replication-blocking lesions with increased DDR foci at telomeres, which are known to accumulate high levels of oxidative damage [51]. Hemizygous mice showed normal life expectancy but higher spontaneous mutations and elevated risk of tumorigenesis, including lymphomas, sarcomas, and adenocarcinomas [52,53,54].
In several cancers (including lung, colorectal, cervical, prostate, bladder, gastric, hepatic, glioblastoma, osteosarcoma, head and neck, ovarian, and breast) high APE1 expression or aberrant cytoplasmic distribution (Table 1) have been associated with tumour aggressiveness, poor prognosis or increased resistance to DNA-damaging agents [55,56]. For instance, in breast cancer, high APE1 expression has been reported in tumor-initiating cells [57], potentially protecting these cells from irradiation-induced oxidative stress and consequent senescence. On the other hand, the presence of cells with low/deficient APE1 expression may be linked to a good prognosis because this increases senescence, which acts as a tumour suppressor. Some somatic mutations have also been found in glioblastoma [58] and endometrial cancer [59], including the R237C substitution, which leads to reduced exonuclease activity [60]. Besides, some polymorphisms in the APE1 promoter have been associated with a decreased risk of lung cancer [61,62].

2.3. ARTEMIS

ARTEMIS, also known as SNM1C/DCLRE1C, is a member of the metallo-b-lactamase superfamily, characterized by their ability to hydrolyze DNA or RNA. ARTEMIS is essential for the NHEJ pathway, where it removes 5′ and 3′-overhangs to join duplex DNA ends or hairpin openings [63,64]. It also facilitates early site-specific chromosome breakage during apoptosis [23]. Although its main nuclease activity acts in a protein kinase C (PKC)-dependent manner, its 5–3′ exonuclease activity is independent of PKC and allows it to function more efficiently in 1- or 2-nucleotide 5′ overhangs, which are too short for endonucleolytic activity and occur following exposure to IR [63,64]. Hence, cells or patients lacking ARTEMIS cannot repair damage caused by IR [65,66] or alkylating agents used in chemotherapy [67]. Moreover, patients with deficiency or mutations (frequently found among Native Americans [68,69]) in ARTEMIS suffer from severe combined immunodeficiency (SCID) (TBNK+) [66] owing to the importance of NHEJ during B and T lymphocyte maturation, where V(D)J recombination is initiated by the creation of DSBs. ARTEMIS null mice also present SCID (Table 2) but they do not exhibit a higher risk of cancer [70]. However, when combined with Trp53 loss, accelerated tumorigenesis has been observed. These mice develop especially aggressive B-cell lymphomas, indicating that ARTEMIS acts as a tumour suppressor in the absence of p53 [71]. Therefore, the defective function of ARTEMIS leads to unrepaired DSBs and malignant transformation of cells that escape apoptosis. ARTEMIS can also act as a negative regulator of p53 in response to oxidative stress induced by mitochondrial respiration. It can also interact with p53 and DNA-PK, inhibiting p53 phosphorylation and activation [72].
Downregulation of ARTEMIS occurs in chronic myeloid leukaemia cell lines, which are characterized by increased levels of reactive oxygen species (ROS) that lead to DNA damage, including DSBs. In these cells, the low levels of the protein cause abnormal processing of DSBs with decreased stability of DNA-PK complexes at DNA ends [73]. Hypomorphic mutations, resulting in truncation of the last exon, have been described in patients with aggressive Epstein-Barr virus-associated B-cell lymphoma (Table 1). Although these patients did not show SCID, they showed low diversity in V(D)J junctions [74,75]. These findings were confirmed in mouse models [76].
The fact that cells deficient in ARTEMIS are more sensitive to radiation has been used as a therapeutic approach. A peptide inhibiting the interaction between ARTEMIS and DNA ligase IV, which is needed for its nuclease activity, has been used as a radiosensitizer that delays DNA repair and synergizes with irradiation to inhibit cell proliferation and induce cell cycle arrest and apoptosis [77].

2.4. EXD2

EXD2 (3′–5′ exonuclease domain-containing protein 2) has a conserved exonuclease domain with high sequence similarity to WRN (explained below). It only functions as an exonuclease when the protein oligomerizes and it can discriminate substrate (DNA or RNA) depending on the metal cofactors [78]. EXD2 localizes at the mitochondrial membrane, where it regulates mitochondrial translation [79], and the nucleus, where it promotes genome stability by acting on replication forks and DSB repair. EXD2 is recruited to replication forks upon replication stress to counteract fork reversal by suppressing the uncontrolled degradation of nascent DNA, allowing efficient fork restart [80]. This protection of the replicating fork is shared with BRCA1/2. Therefore, in the absence of both proteins the unprotected replication forks collapse, resulting in genome instability and compromised cell survival [80]. EXD2 is also essential for the repair of DSBs by HR. It interacts with the MRN (MRE11-RAD50-NBS1) complex to accelerate 3′ resections of double-stranded DNA (dsDNA), both short- and long-range [81]. Cells deficient in EXD2 show spontaneous chromosomal instability and are sensitive to DNA damage induced by anti-cancer agents such as IR and campthotecin [81], thus EXD2 is a good target for the development of a new anti-tumour treatment. So far, no studies have analysed the expression or the presence of mutations in human tumours.

2.5. EXO1

Exonuclease 1 (EXO1) is a member of the Rad2/XPG family, which contains DNA endonuclease, RNase H, and 5′–3′ exonuclease domains [82]. EXO1 (together with FEN1 and POLD) is essential for removing primers and for Okazaki fragment maturation during replication [11,83]. It is also involved in several DDR pathways such as MMR, where it is recruited by MutSα, MutSβ, and MutLα to degrade the newly synthesized DNA containing the replication error [11,83]; and HR, where it resects DNA in DSBs to allow RAD51 loading and strand exchange [84,85,86]. Upon DNA damage, EXO1 is involved in the recruitment of translesion synthesis (TLS) polymerases to sites of UV damage [87] and in the enlargement of single-stranded DNA (ssDNA) gaps to activate the ATR checkpoint by NER [85,88].
EXO1 has been associated with different types of tumours and its overexpression causes an increase in its DNA repair activity and genome instability. Overexpression of EXO1 occurs in prostate [89,90], breast [91,92,93], ovarian (cell lines) [94], lung [95], liver [96,97], bladder [98] and melanoma [99] cancer patients (Table 1). Moreover, mutations in the exonuclease domain resulting in loss of function, such as the A153V and N279S mutations, are found in colorectal and small intestine tumours [100]. These types of tumours also present the E109K mutation, which does not disrupt exonuclease activity, but, as it is localized in the PAR-binding motif, hinders its recruitment to DNA damage sites. In addition, several EXO1 polymorphisms have been associated with a high risk of prostate [101], ovarian [102], lung [103,104,105], oral [106], liver [107], colon [108] and stomach [109] cancer, whereas other variants have shown protective roles in tissues like liver [110] and colon [111].
The effects of EXO1 inactivation (E109K mutation) [85] or deletion (Exo1 knockout (KO)) [112] have been studied in mouse models (Table 2). Both mutant mice showed significantly reduced survival and accelerated tumorigenesis compared to wt mice. However, they showed differences in tumour spectrum. While Exo1 KO predominantly develops lymphomas, mutated mice (Exo1E109K) develop sarcomas and adenomas. The different patterns of tumorigenesis can be attributed to the DSBR deficiency in mutated mice whereas in Exo1 KO mice both the DSBR and MMR pathways are disrupted.
EXO1 activates the immune system in mice with an MLH1-deficient background through the activation of the cGAS-STING pathway [113]. Under normal circumstances, MutLα regulates the activity of EXO1 to generate the appropriate length of ssDNA. However, in the absence of this regulation, EXO1 induces excessive DNA degradation, producing unprotected ssDNA. These events lead to DNA breaks, chromosome abnormalities, and the release of nuclear DNA into the cytoplasm leading to cGAS-STING pathway activation and thus a type I IFN innate immune response. Therefore, it has been proposed that combining radiation and immunotherapy in MLH1-defective patients will be beneficial.

2.6. EXOG

EXOG (Exo/Endonuclease G) is a mitochondrial (mt) endo/5′–3′exonuclease with a preference for ssDNA [114,115]. It forms a complex with the mt repair proteins to remove the 5′-blocking oxidized residues of SSBs in the mt genome by BER. Therefore, depletion of EXOG induces persistent SSBs in the mtDNA, enhances ROS levels, and induces mt dysfunction, triggering the intrinsic apoptotic pathway [116]. This mechanism is especially important in tissues with elevated levels of oxidative agents such as human lung adenocarcinoma tumours, and with high levels of hydrogen sulphide (H2S)-producing enzymes. Elevated levels of H2S stimulate mtDNA repair through sulfhydration of EXOG, which increases its interaction with mt repair proteins to enhance DNA repair [117].
EXOG participates in mtDNA replication. In this process, RNase H1 removes all the RNA primers apart from two nucleotides that remain attached to the 5′end of the nascent DNA. EXOG removes this dinucleotide of the RNA/DNA hybrid duplex, maintaining mitochondrial genome integrity [118]. Since the identification of EXOG in 2008, only one report has associated EXOG with cancer: a missense mutation was found in a familiar case of appendiceal mucinous tumours, an extremely rare disease with uncertain genetic aetiology [119].

2.7. FAN1

FANCD2/FANCI-associated nuclease 1 (FAN1) is a 5′ flap structure-specific endonuclease and 5′–3′ exonuclease with broad substrate specificity [12]. It is essential to maintain chromosomal stability and resolve ICLs. Although its exact mechanism of action remains unclear, it is thought that FAN1 makes 2–6 nucleotide incisions at the sides flanking the ICLs, generating a suitable substrate for other nucleases and polymerases. In addition, it can participate in MMR, interacting with MutLα in the absence of EXO1, or cleave D-loop structures formed during HR. In response to replication stress, FAN1 also controls the progression of stalled replication forks, where it is recruited by Ub-FANCD2 (Fanconi anemia pathway) [12].
A deficiency of FAN1 in humans leads to chromosomal abnormalities (caused by failure of the replication fork) that can cause rare kidney and neurological diseases such as schizophrenia, epilepsy, and autism [12,120]. Although the loss of FAN1 function does not increase the burden of cancer [121], some mutations have been found in tumours, including mutations abolishing nuclease/exonuclease activity. For example, the p.M50R mutation occurs as a germline mutation in hereditary pancreatic cancer [122] and it also increases the risk of colorectal cancer [123] (Table 1). Additional germline mutations have been suggested to increase susceptibility to colorectal cancer [124] and primary hepatic mucoepidermoid carcinoma [125]. Moreover, mice defective in the nuclease domain (Table 2) develop carcinomas and lymphomas [126].
Loss of FAN1 leads to sensitivity to crosslinking agents, especially in BRCA2-deficient cells [127]. Increased FAN1 expression in tumours refractory to treatment has been observed in breast and ovarian cancers [128]. Therefore, inhibition of FAN1 could be used to sensitize cancer cells to conventional chemotherapy. Additionally, FAN1 functional status in cancer cells might be used as a biomarker to predict response to treatment.

2.8. FEN1

Flap endonuclease 1 (FEN1), also known as DNase IV, belongs to the RAD2 family and is involved in multiple functions via different catalytic activities [13]. FEN (flap-specific endonuclease) activity is responsible for RNA primer removal in the maturation of Okazaki fragments during DNA replication and repairing DNA lesions that have an oxidatively damaged sugar moiety in a PCNA-dependent BER pathway called long-patch BER. EXO (5′ exonuclease) and GEN (gap-endonuclease) activities are important for the resolution of trinucleotide repeat sequence-derived DNA hairpin structures, oligonucleaosomal fragmentation of chromosomes in apoptotic cells, and the resolution of stalled replication forks caused by exogenous insults. In this case, FEN1 forms a complex with WRN to arrest the replication fork and resolve the chicken foot structure or cleave the fork to start the break-induced recombination. These multiple functions are regulated protein-protein interactions, post-translation modifications, and cellular compartmentalization, for example, FEN1 translocates to the nucleus upon DNA damage [13].
FEN1 somatic mutations have been found in non-small cell lung carcinoma, melanoma, and oesophageal cancers, some of them inactivating its exonuclease activity [129]. To study the role of FEN1 in cancer, mouse models have been developed (Table 2). Fen1+/− mice (Fen1 KO is embryonically lethal [130]) have an increased risk of tumour development, especially lymphomas [131], and tumorigenesis is further increased in combination with other alterations such as Apc1638N; these mice present reduced survival and increased intestinal adenocarcinomas compared to Apc1638N alone [131]. Mice expressing the FEN1 E160D mutation (abrogates the EXO and GEN activities but not FEN activity), which leads to spontaneous mutations and the accumulation of incompletely digested DNA fragments in apoptotic cells [129], developed autoimmunity, chronic inflammation, and lung, testis/ovary, liver, kidney, spleen, stomach and lymphoma cancers. This phenotype is related to higher spontaneous mutation rates and the accumulation of apoptotic DNA in mutated cells leading to the DNA damage response and inflammation. Another example is the L209P mutation, found in colorectal cancer patients [132]. This mutated protein has lost all three activities and acts as a dominant-negative isoform. Mutated cells show high sensitivity to DNA damage, which causes genomic instability and transformation.
FEN1 is expressed in proliferating cells and is overexpressed in different tumours such as prostate [133], testis [134], lung [134,135], brain [134], gastric [136] and breast [137,138] (Table 1). In some cases, its overexpression is correlated with hypomethylation of the FEN1 promoter and linked to increased tumour grade and aggressiveness [136,137]. FEN1 polymorphisms have been associated with an elevated risk of lung, ovary, bladder, breast, glioma, and digestive cancers. In contrast, a protective role was attributed to some other variants in oesophagus, breast, and leukaemia cancers [139,140,141,142,143,144,145,146,147,148,149,150,151,152].

2.9. MRE11A

MRE11A is an ssDNA endonuclease/dsDNA 3′–5′ exonuclease of the MRN complex that is involved in DNA repair (HR and alternative NHEJ) following DSBs lesions, meiotic recombination, cell cycle checkpoints, and maintenance of telomeres. Its exonuclease activity plays an essential role in DDR, degrading DNA between the endonucleolytic incision sites, which creates an entry site for the long-range resection nucleases [153]. Mutations in MRE11A have been found in some types of cancer characterized by chromosomal instabilities such as breast, endometrium, and colon [154,155,156,157]. Mutations Y187C and H52S inactivate MRE11A exonuclease but not endonuclease activity [158]. Some frameshift mutations generate splicing variants that lead to exon loss. HCT116 cells (colon cancer cell line) have a mutant protein without exons 5–7, where the exonuclease domain is located, leading to the accumulation of unrepaired DNA [156]. MRE11A has also demonstrated potential as a predictive marker for radiotherapy in bladder cancer patients (Table 1), where high expression of MRE11A has been associated with a good prognosis [159].

2.10. p53

p53 is known to be the “guardian of the genome”, ensuring genetic stability through several roles that include control of the cell cycle, senescence, apoptosis, and DNA repair in response to oncogene activation, DNA damage, and other stress signals. Although p53 acts mainly as a transcriptional factor, it has 3′–5′ exonuclease activity in its core domain, where the sequence-specific DNA binding domain is located [160,161,162,163]. These two activities are mutually exclusive; therefore, exonuclease activity is mainly cytoplasmic [164,165,166]. It has been suggested that its exonuclease function is its most ancient function since this domain is present even in the ancestral p53 in invertebrates [167]. p53 shows a preference for ssDNA (although it can also process dsDNA, single-stranded RNA (ssRNA), and double-stranded RNA (dsRNA)), can remove 3′-terminal miss-pairs, and has a proofreading function when interacting with exonuclease-deficient polymerases [164,168,169]. Exonuclease activity has been observed in unstressed cells, but it also responds to exogenous stimuli such as DNA-damaging agents. In this scenario, exonuclease activity is not involved in cell cycle arrest but is essential for the induction of apoptosis in DNA-damaged cells [165]. Several core domain mutations have been found in cancer. However, their specific effect on exonuclease activity has not been assessed.

2.11. PLD3 and PLD4

PLD3 and PLD4 belong to the phospholipase D (PLD) family [170] and are characterized by their HKD motifs. They are glycosylated transmembrane proteins located in endolysosomes and surprisingly, in contrast to their family members PLD1/2, do not possess phospholipase activity. Instead, their different HKD amino acid composition allows them to process ssDNA from 5′ to 3′ [26], degrading endogenous ssDNA and thereby preventing autoimmune pathologies like rheumatoid arthritis [29]. In addition, both can degrade ssRNA, thus limiting autoinflammation triggered by both endosomal TLR and cytoplasmic STING nucleic acid sensing pathways [27].
PLD family members have a well-established role in promoting tumorigenesis in multiple types of cancers [171]. Although PLD4 and PLD3 have been linked through murine models (Table 2) and genome-wide association studies to autoinflammatory diseases [172] and late-onset Alzheimer’s disease [173], respectively, very little evidence exists regarding their involvement in cancer. PLD3 has been associated with a favourable prognosis in osteosarcoma and included in a prognostic gene signature of immune cell infiltration [174,175] (Table 1). Likewise, PLD4 expression has been proposed to predict, together with other genes, better survival of HER2-positive breast cancer patients [176] and, when expressed in M1 macrophages, it might have antitumoral effects in colon cancer [177]. Overall, our knowledge of the role of PLD3/4 in cancer is limited, and thus additional research is warranted.

2.12. POLD and POLE

Polymerase delta (POLD) and polymerase epsilon (POLE) contain 3′–5′ exonuclease proofreading domains. Several studies have suggested that this intrinsic proofreading exonuclease activity plays a critical role in suppressing carcinogenesis. Increased epithelial cancer was observed in mice deficient for Pold proofreading exonuclease (Table 2) (Pold1D400A/D400A) [178]. Moreover, germline or somatic mutations in the exonuclease domain of POLE were found to increase the mutation rate and risk of cancer development in the colon and endometrium [179]. Although somatic POLD exonuclease domain mutations are less frequent, they were observed in colon, endometrium, and melanoma cancers [180,181] (Table 1). Patients with these mutations have an excellent prognosis and respond well to immunotherapy because the high mutation rate increases the probability of neoantigens, which are recognized by the immune system [179,181].

2.13. RAD9A

RAD9A is a multifunctional protein that contains a 3′–5′ exonuclease domain in its N-terminal portion [182], although the exact function of this activity remains unknown. This protein is involved in several cellular functions such as apoptosis (it contains a BH3 domain), but its main role is to control the cell cycle checkpoint and DNA damage repair as an early DNA damage sensor of SSBs and DSBs [182]. RAD9A participates in multiple repair pathways such as BER (interacting and activating APE1), MMR (interacting with MLH1, MSH2, MSH3, and MSH6), ICL (activating FANCD2 through ATR activation), and HR (interacting with RAD51) [182]. Consequently, cells deficient in RAD9A show spontaneous chromosome aberrations and are more sensitive to DNA-damaging agents such as hydroxyurea, UV light, and IR [183].
Owing to its involvement in multiple and varied cellular functions, RAD9A has a dual role in cancer, acting as a tumour suppressor or promoter depending on the tissue. RAD9A is overexpressed and accumulated in the nucleus of samples from non-small cell lung carcinomas, and is correlated with high proliferation [184,185]. Overexpression of RAD9 was also observed in thyroid [186], prostate [187,188], and breast cancer [189] (Table 1). In breast and prostate cancer, overexpression is due to amplification of the 11q13 chromosome (where the RAD9 gene is located) or differential intron methylation in the RAD9 gene. The introns contain sequences that inhibit RAD9 expression but are suppressed upon methylation, an event that occurs in childhood leukaemia [190]. Increased levels of RAD9A were correlated with bigger tumours, local recurrence, and higher aggressiveness. However, other types of tumours like gastric carcinomas showed decreased expression [191]. The same could be true for skin cancers, since the skin conditional Rad9-deficient mouse (total KO is lethal [192]) showed enhanced tumour development upon application of carcinogen [193] (Table 2).

2.14. TREX1

Three-prime repair exonuclease 1 (TREX1 or DNase III) is a non-processive 3′–5′ exonuclease that degrades ssDNA and dsDNA from the 3′-ends [194,195]. TREX1 is ubiquitously expressed and localized in the endoplasmic reticulum membrane. It plays a major role in DNA-driven immune responses, where it is involved in self- and non-self-DNA degradation, limiting the activation of DNA-sensing and -signalling pathways, such as the cGAS-STING pathway. Thus, it prevents aberrant interferon-mediated inflammatory responses and autoimmunity [196,197,198,199,200]. Furthermore, TREX1 has been implicated in DNA degradation during granzyme A-mediated cell death [21].
Independently of its exonuclease activity, TREX1 reduces glycan-driven immune responses by interacting with the oligosaccharyltransferase complex, contributing to its correct function [201]. Consistently, loss of function mutations have been associated with inflammatory and autoimmune diseases. Of note, mutations located in the exonuclease domains (N-terminus region) are mostly linked to Aicardi-Goutières syndrome and systemic lupus erythematosus. Mutations in the endoplasmic reticulum localization and oligosaccharyltransferase interaction domain (C-terminus region) are mainly linked to retinal vasculopathy with cerebral leukodystrophy [202,203,204,205]. In agreement with the major role of TREX1 as an anti-inflammatory player, Trex1 knockout mice [206] and TREX1D18N exonuclease defective mice [207,208] develop an inflammatory systemic phenotype, but they are not tumour prone (Table 2). Nevertheless, TREX1 can influence genomic stability and DDR in distinct ways. For instance, TREX1 is induced after DNA damage, favouring DNA repair [19] and interacting and stabilizing PARP1 [209]. Moreover, TREX1-deficient cells exhibited increased levels of p53 and p21 and ATM-dependent checkpoint activation [210], which suggests chronic activation of DDR. However, TREX1 can drive chromosome mis-segregation and error-prone DNA repair in tumoral cells undergoing telomere crisis, thus fostering genomic instability [211]. Finally, TREX1 locates to micronuclei, degrading DNA when their membranes break, preventing its cytoplasmic sensing and supporting the chromosomal instability of tumours [197].
Depending on the type of tumour, TREX1 expression is upregulated or downregulated (Table 1). TREX1 overexpression is observed in oesophageal [212] and cervical [213] cancers but is downregulated in melanoma [214] and osteosarcoma [215], where TREX1 expression is only observed in non-metastatic patients. Conflicting results have been found in breast cancer, with some studies showing overexpression and some downregulation [216,217]. Recently, a potentially pathogenic TREX1 variant was found in a small cohort of familial colorectal cancer type X (FCCTX), although the functional consequences of the variant were not assessed [218]. Of note, TREX1 is induced in carcinoma cells by irradiation [199,219] and in glioma [214], melanoma [220], and nasopharyngeal [221] cells by anticancer drugs, triggering a pro-survival response. Because preventing TREX1 from degrading accumulated cytosolic DNA renders the cGAS-STING pathway active with the consequent production of type I IFNs, TREX1 has attracted great interest as a target to elicit antitumour immunity [222].

2.15. TREX2

Three-prime repair exonuclease 2 (TREX2) is a 3′–5′ exonuclease that is highly homologous to TREX1 [223,224], sharing similar biochemical and structural features [225]. However, in contrast to TREX1, TREX2 shows restricted expression in keratinocytes, localizes in nuclei, and plays a relevant pro-inflammatory role in keratinocytes [22,224], [226] without activating DNA-driven immune responses [227]. TREX2 facilitates nuclear DNA degradation in stressed keratinocytes, thus promoting cell death [22,224,226,227]. Interestingly, TREX2 has been shown to improve targeted CRISPR/Cas9 efficiency [228,229,230] by increasing deletion size and preventing perfect DNA repair, thereby avoiding repeated Cas9 cleavage and chromosomal translocations [231].
Studies on the role of TREX2 as a tumour suppressor have reported contrasting results. In artificial settings, chromosomal and genomic instability were reported using Trex2-null and mutated embryonic stem cells [232,233]; however, TREX2 deficiency in mice (Table 2) does not lead to a tumour prone phenotype [22,224]. Depending on the DNA repair status of embryonic stem cells, TREX2 may display a dual function in the DDR pathway, dependent and independent of its exonuclease activity, facilitating or avoiding replication fork instability and mutations [234,235]. However, neither genomic nor chromosomal instability are observed in cells from Trex2 knockout mice or in keratinocytes, where Trex2 is highly expressed, nor in embryonic stem cells, in which TREX 2 expression is not detectable [224]. Nevertheless, Trex2 knockout mice show increased susceptibility to DNA damage-induced skin tumorigenesis. TREX2 interacts with phosphorylated H2AX histone, which is a critical player in both DNA repair and cell death and is recruited to low-density nuclear chromatin and micronuclei. Upon DNA damage, TREX2 participates in DNA repair but mostly contributes to DNA degradation of fragmented DNA, promoting cell death of damaged keratinocytes and favouring an antitumoral immune response, supporting a major role of TREX2 as a proapoptotic tumour suppressor in keratinocyte-driven tumours [22,224].
TREX2 deregulation and genetic alterations in cancer mostly indicate the role of TREX2 as a tumour suppressor. In cutaneous squamous cell carcinomas (cSCCs) and head and neck SCCs (HNSCCs) high expression of TREX2 was found in well-differentiated tumours while its expression was not detected in metastatic samples [22] (Table 1). Moreover, epigenetic regulation of TREX2 was observed in colorectal and laryngeal cancer [236]. Reduced DNA methylation in the TREX2 intragenic gene locus is associated with elevated expression and better overall survival of patients. In contrast, TREX2 missense mutations and upregulation in colorectal cancer have been associated with reduced survival [237]. In this regard, SNPs in TREX2 are more frequent in patients with head and neck SCCs than in healthy individuals [22].

2.16. WRN

Werner Syndrome protein (WRN) (also known as RECQL2) is a multifunctional protein that contains four functional domains, including a 3′–5′ exonuclease domain that can recognize a variety of DNA substrates [238]. This protein is crucial for genome stability through its involvement in DNA replication, recombination, and repair, although the specific relevance of the exonuclease domain has not yet been determined. Germline mutations in WRN lead to Werner Syndrome, characterized by premature aging and higher susceptibility to a broad spectrum of epithelial and mesenchymal tumours like sarcomas, melanoma, thyroid, and breast cancer [238]. Although no somatic mutations have been described in sporadic tumours, its expression is downregulated due to epigenetic inactivation or loss of heterozygosity in several solid tumours such as colorectal cancer and breast tumours [239,240]. Low expression of nuclear WRN has been associated with a worse prognosis and promoter hypermethylation as a predictor of good clinical response to DNA-damaging treatments [239,240,241,242].
Table 1. Exonuclease alterations in cancer.
Table 1. Exonuclease alterations in cancer.
GeneAlterationType of CancerBiomarkerRef.
AENHigh expressionProstateHigh-risk recurrence[40]
ColorectalReduced survival[41]
APE1Exonuclease mutationsGlioblastoma, endometrial [59,60]
High expressionLung, colorectal, cervical, prostate, bladder, gastric, hepatic, glioblastoma, osteosarcoma, head, and neck, ovarian, breastTumour aggressiveness, poor prognosis[55,56]
ARTEMISHypomorphic mutationsLymphomaHigh risk[74,75]
EXO1Exonuclease inactivating mutationsColorectal tumours, small intestine tumours [100]
High expressionProstate, breast, lung, liver, bladder, melanoma [89,90,91,92,93,95,96,97,98,99]
FAN1Exonuclease inactivating mutationsPancreatic, colorectal, hepaticHigh risk[122,123,124,125]
High expressionOvarianPoor prognosis[128]
FEN1High expressionProstate, testis, lung, brain, gastric, breastIncreased tumour grade and aggressiveness[133,134,135,136,137,138]
SNPLung, ovary, bladder, breast, glioma, digestiveHigh risk[139,140,141,142,143,144,145,146,147,148,149,150,151,152]
Esophagus, breast, leukemiaProtective role[144,146,147]
MRE11AExonuclease inactivating mutationsBreast, endometrium, colon [154,155,156,157]
High expressionBladderGood prognosis[159]
PLD3High expressionOsteosarcomaGood prognosis[174,175]
PLD4High expressionHER2-positive breast cancerBetter survival[176]
POLDSomatic exonuclease domain mutationsColon, endometrium, and melanomaGood prognosis[179,180,181]
POLEExo domains mutatedColon, endometriumHigh risk and increased mutation rate[179]
RAD9High expressionLung, thyroid, prostate, breastBigger tumours, recurrence, and aggressiveness[184,185,186,187,188,189]
Low expressionGastric [191]
TREX1High expressionEsophageal, cervix [212,213]
Low expressionMelanoma, osteosarcoma [214,215]
TREX2High expressionLow-grade HNSCC, laryngealGood prognosis [22,236]
ColorectalReduced survival[237]
Low expressionMetastatic HNSCC [22]
WRNSomatic mutationsSarcomas, melanoma, thyroid, breast [238]
Low expressionColorectal, breastBad prognosis[239,240]
Table 2. Murine strains modelling exonuclease gene functions.
Table 2. Murine strains modelling exonuclease gene functions.
Exonucl.Mutant MiceAlterationPhenotypeRef.
APE1Ape1−/−Gene deletionLethal[50]
Ape1+/−HemizygousCancer prone, lymphomas, sarcomas & adenocarcinomas[52,53,54]
ARTEMISArtN/NGene deletionSevere combined immunodeficiency[70]
ArtN/NTrp53N/NGene deletionIncreased carcinogenesis in Art vs. p53 null mice[72]
EXO1Exo−/−Gene deletionLymphoma; reduced survival; sterility[112]
FAN1Fannd/ndNuclease defectiveCancer prone, carcinomas & lymphomas[126]
FEN1Fen1−/−Gene deletionLethal[130]
Fen1+/−HemizygousTumours, mainly lymphomas[131]
Fen1+/− Apc1368NHemizygous MutationIncreased adenocarcinomas & decreased survival compared to Apc1268N[131]
Fen1E160DInactivation of exo- & endonuclease activitiesAutoimmunity, chronic inflammation, and tumours. Spontaneous mutations; accumulation of non-digested DNA in apoptotic cells.[129]
PLD3 and PLD4Pld3−/−Gene deletionNo inflammation[26]
Pld4−/−Gene deletionInflammation, splenomegaly, high IFNγ levels[26]
Pld3−/−Pld4−/−Gene deletionLethal liver inflammation, hemophagocytic lymphohistiocytosis, high IFNγ levels[27]
POLDPold1D400AExonuclease domain mutatedIncreased epithelial cancer[178]
RAD9Rad9−/−Gene deletionLethal[192]
Rad9K5−/−Gene deletion in keratinocytesEnhanced tumour development upon exposure to carcinogen[193]
TREX1Trex1−/−Gene deletionNot cancer-prone. Inflammatory myocarditis[206]
Trex1D18NExonuclease defectiveNot cancer-prone. Systemic inflammation. Lupus-like inflammatory syndrome.[208]
TREX2Trex2−/−Gene deletionNot cancer-prone. Increased carcinogenesis upon exposure to genotoxins. Reduced inflammation.[22,224,226]

3. Outlook

The DDR encompasses different sensor and effector proteins, including exonucleases, that work together with the final aim of limiting damage, both for the cell and the organism. Almost all living organisms have evolved to possess a battery of mechanisms to ensure the preservation of their hereditary material, demonstrating the ubiquitous urgency to respond to damaged DNA [243,244]. DNA exonuclease activity is required for basic cell processes, such as the synthesis and repair of damaged DNA, cell death, and inflammation (Figure 1) which are important to maintain homeostasis and prevent diseases and cancer. While not all DNA exonucleases are directly involved in the DDR, their actions can indirectly alter it. The functional interplay between the exonucleases described in this revision is surprisingly significant (Figure 2) considering that some of them are highly confined and spatially separated in different tissues and compartments of the cell. While most exonucleases reside in the nucleus, EXOG and EXD2 are specifically located in the mitochondria, and TREX1, PLD3, and PLD4 in the cytosol. Additionally, others have fluctuating levels depending on the cell cycle (APEX1) or are only expressed in keratinocytes (TREX2). Either interacting in multiprotein complexes or by their own action, exonucleases contribute in different manners to execute a proper DDR.
DNA exonucleases might be required for the survival of some cancer cells that acquire dependency on normal cellular functions, such as DNA repair, and thus, produce a non-oncogene addiction [245]. DNA exonucleases may favour the appearance of mutations in healthy cells, converting them into tumour cells, or may be necessary for certain tumours to keep the inherent genomic instability of malignancy under control, aiding tumour cells to avoid cell death, or impact on the immune response. In this manner, aberrant DNAse activity in tumours have been suggested to be exploited as a molecular whistleblower for diagnosis [246]. Hence, these types of enzymes are promising targets for cancer treatment promoting synthetic lethality and early detection of certain types of cancer. For instance, blocking the function of those exonucleases that degrade cytoplasmic DNA preventing inflammation, such as TREX1, PLD3 and PLD4, would activate DNA sensors producing an IFN response and leading to antitumor immunity.
Exonuclease genetic alterations, changes in their activity and abnormal expression in human tumours together with functional studies with murine models point to a clear contribution of these proteins to cancer onset and development (Table 1 and Table 2). It is puzzling how, depending on the type of cancer, some exonucleases are overexpressed or downregulated. It is worth noting the different tissue-specific metabolic requirements and byproducts as well as the differences in protein expression that might influence the types of genes being expressed. Moreover, some types of cancers are characterized by a particularly elevated genomic instability and this itself could influence the expression of the transcription factors required for exonuclease expression. Nonetheless, each individual tumour is different and the genetic background of the individual as well as environmental factors such as diet, temperature, medications, or immune cell infiltration can influence gene expression, and result in these not surprising differences. On the other hand, the existing knowledge on some exonucleases, such as AEN, EXD2, EXOG, PDL3/4, TREX1, and TREX2, is relatively poor and sometimes conflicting. Advance in the understanding of the mechanisms and functions of each specific exonucleases would strengthen their value as potential targets for cancer treatment and/or as biomarkers.

Author Contributions

Introduction (J.M., C.S.); Exonucleases and Cancer (L.M., J.M., C.S.); Outlook (J.M., C.S.). Figure 1 (L.M., C.S.); Figure 2 (J.M.); Table 1 and Table 2 (L.M., J.M., C.S.). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by research grant PID2020-114477RB-I00 from the MCIN/ AEI /10.13039/501100011033 to CS.

Data Availability Statement

Figure 1 was created with BioRender.com. The functional association between exonucleases was assessed and graphically represented using String: https://string-db.org/.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Watson, J.D.; Crick, F.H.C. Molecular structure of nucleic acids: A structure for deoxyribose nucleic acid. Nature 1953, 171, 737–738. [Google Scholar] [CrossRef]
  2. Westheimer, F.H. Why nature chose phosphates. Science 1987, 235, 1173–1178. [Google Scholar] [CrossRef] [PubMed]
  3. Zuo, Y.; Deutscher, M.P. Exoribonuclease superfamilies: Structural analysis and phylogenetic distribution. Nucleic Acids Res. 2001, 29, 1017–1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Baranovskii, A.G.; Buneva, V.N.; Nevinsky, G.A. Human deoxyribonucleases. Biochem. Biokhimiia 2004, 69, 587–601. [Google Scholar] [CrossRef] [PubMed]
  5. Dehé, P.-M.; Gaillard, P.-H.L. Control of structure-specific endonucleases to maintain genome stability. Nat. Rev. Mol. Cell Biol. 2017, 18, 315–330. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, W. Nucleases: Diversity of structure, function and mechanism. Q. Rev. Biophys. 2010, 44, 1–93. [Google Scholar] [CrossRef] [PubMed]
  7. Zhou, B.-B.S.; Elledge, S.J. The DNA damage response: Putting checkpoints in perspective. Nature 2000, 408, 433–439. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Kastenhuber, E.R.; Lowe, S.W. Putting p53 in context. Cell 2017, 170, 1062–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Liu, T.-C.; Guo, K.-W.; Chu, J.-W.; Hsiao, Y.-Y. Understanding APE1 cellular functions by the structural preference of exonuclease activities. Comput. Struct. Biotechnol. J. 2021, 19, 3682–3691. [Google Scholar] [CrossRef]
  11. Keijzers, G.; Liu, D.; Rasmussen, L.J. Exonuclease 1 and its versatile roles in DNA repair. Crit. Rev. Biochem. Mol. Biol. 2016, 51, 440–451. [Google Scholar] [CrossRef]
  12. Deshmukh, A.L.; Porro, A.; Mohiuddin, M.; Lanni, S.; Panigrahi, G.B.; Caron, M.-C.; Masson, J.-Y.; Sartori, A.A.; Pearson, C.E. FAN1, a DNA repair nuclease, as a modifier of repeat expansion disorders. J. Huntingt. Dis. 2021, 10, 95–122. [Google Scholar] [CrossRef] [PubMed]
  13. Zheng, L.; Jia, J.; Finger, L.D.; Guo, Z.; Zer, C.; Shen, B. Functional regulation of FEN1 nuclease and its link to cancer. Nucleic Acids Res. 2010, 39, 781–794. [Google Scholar] [CrossRef] [Green Version]
  14. Nishino, T.; Morikawa, K. Structure and function of nucleases in DNA repair: Shape, grip and blade of the DNA scissors. Oncogene 2002, 21, 9022–9032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Collins, J.A.; Schandl, C.A.; Young, K.K.; Vesely, J.; Willingham, M.C. Major DNA fragmentation is a late event in apoptosis. J. Histochem. Cytochem. 1997, 45, 923–934. [Google Scholar] [CrossRef]
  16. Bosurgi, L.; Hughes, L.D.; Rothlin, C.V.; Ghosh, S. Death begets a new beginning. Immunol. Rev. 2017, 280, 8–25. [Google Scholar] [CrossRef] [PubMed]
  17. Lee, J.-H.; Koh, Y.A.; Cho, C.-K.; Lee, S.-J.; Lee, Y.-S.; Bae, S. Identification of a novel ionizing radiation-induced nuclease, AEN, and its functional characterization in apoptosis. Biochem. Biophys. Res. Commun. 2005, 337, 39–47. [Google Scholar] [CrossRef] [PubMed]
  18. Kawase, T.; Ichikawa, H.; Ohta, T.; Nozaki, N.; Tashiro, F.; Ohki, R.; Taya, Y. p53 target gene AEN is a nuclear exonuclease required for p53-dependent apoptosis. Oncogene 2008, 27, 3797–3810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Christmann, M.; Tomicic, M.T.; Aasland, D.; Berdelle, N.; Kaina, B. Three prime exonuclease I (TREX1) is Fos/AP-1 regulated by genotoxic stress and protects against ultraviolet light and benzo(a)pyrene-induced DNA damage. Nucleic Acids Res. 2010, 38, 6418–6432. [Google Scholar] [CrossRef] [PubMed]
  20. Lieberman, J. Granzyme A activates another way to die. Immunol. Rev. 2010, 235, 93–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Chowdhury, D.; Beresford, P.J.; Zhu, P.; Zhang, D.; Sung, J.-S.; Demple, B.; Perrino, F.W.; Lieberman, J. The exonuclease TREX1 is in the SET complex and acts in concert with NM23-H1 to degrade DNA during granzyme a-mediated cell death. Mol. Cell 2006, 23, 133–142. [Google Scholar] [CrossRef] [PubMed]
  22. Manils, J.; Gómez, D.; Salla-Martret, M.; Fischer, H.; Fye, J.M.; Marzo, E.; Marruecos, L.; Serrano, I.; Salgado, R.; Rodrigo, J.P.; et al. Multifaceted role of TREX2 in the skin defense against UV-induced skin carcinogenesis. Oncotarget 2015, 6, 22375–22396. [Google Scholar] [CrossRef] [PubMed]
  23. Britton, S.; Frit, P.; Biard, D.; Salles, B.; Calsou, P. ARTEMIS nuclease facilitates apoptotic chromatin cleavage. Cancer Res. 2009, 69, 8120–8126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Santa, P.; Garreau, A.; Serpas, L.; Ferriere, A.; Blanco, P.; Soni, C.; Sisirak, V. The role of nucleases and nucleic acid editing enzymes in the regulation of self-nucleic acid sensing. Front. Immunol. 2021, 12, 629922. [Google Scholar] [CrossRef]
  25. Bartok, E.; Hartmann, G. Immune sensing mechanisms that discriminate self from altered self and foreign nucleic acids. Immunity 2020, 53, 54–77. [Google Scholar] [CrossRef]
  26. Gavin, A.L.; Huang, D.; Huber, C.; Mårtensson, A.; Tardif, V.; Skog, P.D.; Blane, T.R.; Thinnes, T.C.; Osborn, K.; Chong, H.S.; et al. PLD3 and PLD4 are single-stranded acid exonucleases that regulate endosomal nucleic-acid sensing. Nat. Immunol. 2018, 19, 942–953. [Google Scholar] [CrossRef]
  27. Gavin, A.L.; Huang, D.; Blane, T.R.; Thinnes, T.C.; Murakami, Y.; Fukui, R.; Miyake, K.; Nemazee, D. Cleavage of DNA and RNA by PLD3 and PLD4 limits autoinflammatory triggering by multiple sensors. Nat. Commun. 2021, 12, 5874. [Google Scholar] [CrossRef]
  28. Rice, G.; Newman, W.G.; Dean, J.; Patrick, T.; Parmar, R.; Flintoff, K.; Robins, P.; Harvey, S.; Hollis, T.; O’Hara, A.; et al. Heterozygous mutations in TREX1 cause familial chilblain lupus and dominant Aicardi-Goutières syndrome. Am. J. Hum. Genet. 2007, 80, 811–815. [Google Scholar] [CrossRef]
  29. Chen, W.-C.; Wang, W.-C.; Okada, Y.; Chang, W.-C.; Chou, Y.-H.; Chang, H.-H.; Huang, J.-D.; Chen, D.-Y. rs2841277 (PLD4) is associated with susceptibility and rs4672495 is associated with disease activity in rheumatoid arthritis. Oncotarget 2017, 8, 64180–64190. [Google Scholar] [CrossRef] [Green Version]
  30. Song, L.; Wang, Y.; Zhang, J.; Song, N.; Xu, X.; Lu, Y. The risks of cancer development in systemic lupus erythematosus (SLE) patients: A systematic review and meta-analysis. Arthritis Res. Ther. 2018, 20, 270. [Google Scholar] [CrossRef] [Green Version]
  31. Hemminki, K.; Liu, X.; Ji, J.; Sundquist, J.; Sundquist, K. Effect of autoimmune diseases on risk and survival in histology-specific lung cancer. Eur. Respir. J. 2012, 40, 1489–1495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. Sperka, T.; Wang, J.; Rudolph, K.L. DNA damage checkpoints in stem cells, ageing and cancer. Nat. Rev. Mol. Cell Biol. 2012, 13, 579–590. [Google Scholar] [CrossRef] [PubMed]
  34. Syeda, A.H.; Hawkins, M.; McGlynn, P. Recombination and replication. Cold Spring Harb Perspect. Biol. 2014, 6, 1–14. [Google Scholar] [CrossRef] [Green Version]
  35. Knijnenburg, T.A.; Wang, L.; Zimmermann, M.T.; Chambwe, N.; Gao, G.F.; Cherniack, A.D.; Fan, H.; Shen, H.; Way, G.P.; Greene, C.S.; et al. Genomic and molecular landscape of DNA damage repair deficiency across the cancer genome atlas. Cell Rep. 2018, 23, 239–254.e6. [Google Scholar] [CrossRef] [Green Version]
  36. Eby, K.G.; Rosenbluth, J.M.; Mays, D.J.; Marshall, C.B.; Barton, C.E.; Sinha, S.; Johnson, K.N.; Tang, L.; Pietenpol, J.A. ISG20L1 is a p53 family target gene that modulates genotoxic stress-induced autophagy. Mol. Cancer 2010, 9, 95. [Google Scholar] [CrossRef] [Green Version]
  37. Kaatsch, H.L.; Becker, B.V.; Schüle, S.; Ostheim, P.; Nestler, K.; Jakobi, J.; Schäfer, B.; Hantke, T.; Brockmann, M.A.; Abend, M.; et al. Gene expression changes and DNA damage after ex vivo exposure of peripheral blood cells to various CT photon spectra. Sci. Rep. 2021, 11, 12060. [Google Scholar] [CrossRef]
  38. Moschella, F.; Torelli, G.F.; Valentini, M.; Urbani, F.; Buccione, C.; Petrucci, M.T.; Natalino, F.; Belardelli, F.; Foà, R.; Proietti, E. Cyclophosphamide induces a type I interferon–associated sterile inflammatory response signature in cancer patients’ blood cells: Implications for cancer chemoimmunotherapy. Clin. Cancer Res. 2013, 19, 4249–4261. [Google Scholar] [CrossRef] [Green Version]
  39. Wu, S.-H.; Hsiao, Y.-T.; Chen, J.-C.; Lin, J.-H.; Hsu, S.-C.; Hsia, T.-C.; Yang, S.-T.; Hsu, W.-H.; Chung, J.-G. Bufalin alters gene expressions associated DNA damage, cell cycle, and apoptosis in human lung cancer NCI-H460 cells in vitro. Molecules 2014, 19, 6047–6057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Long, G.; Ouyang, W.; Zhang, Y.; Sun, G.; Gan, J.; Hu, Z.; Li, H. Identification of a DNA repair gene signature and establishment of a prognostic nomogram predicting biochemical-recurrence-free survival of prostate cancer. Front. Mol. Biosci. 2021, 8, 608369. [Google Scholar] [CrossRef]
  41. Miao, Y.; Zhang, H.; Su, B.; Wang, J.; Quan, W.; Li, Q.; Mi, D. Construction and validation of an RNA-binding protein-associated prognostic model for colorectal cancer. PeerJ 2021, 9, e11219. [Google Scholar] [CrossRef] [PubMed]
  42. Dyrkheeva, N.S.; Khodyreva, S.N.; Lavrik, O.I. Multifunctional human apurinic/apyrimidinic endonuclease 1: Role of additional functions. Mol. Biol. 2007, 41, 402–416. [Google Scholar] [CrossRef]
  43. Kaur, G.; Cholia, R.P.; Mantha, A.K.; Kumar, R. DNA repair and redox activities and inhibitors of apurinic/apyrimidinic endonuclease 1/redox effector factor 1 (APE1/Ref-1): A comparative analysis and their scope and limitations toward anticancer drug development. J. Med. Chem. 2014, 57, 10241–10256. [Google Scholar] [CrossRef] [PubMed]
  44. Thakur, S.; Sarkar, B.; Cholia, R.P.; Gautam, N.; Dhiman, M.; Mantha, A.K. APE1/Ref-1 as an emerging therapeutic target for various human diseases: Phytochemical modulation of its functions. Exp. Mol. Med. 2014, 46, e106. [Google Scholar] [CrossRef]
  45. Laev, S.S.; Salakhutdinov, N.F.; Lavrik, O.I. Inhibitors of nuclease and redox activity of apurinic/apyrimidinic endonuclease 1/redox effector factor 1 (APE1/Ref-1). Bioorg. Med. Chem. 2017, 25, 2531–2544. [Google Scholar] [CrossRef]
  46. Ma, X.; Dang, C.; Min, W.; Diao, Y.; Hui, W.; Wang, X.; Dai, Z.; Wang, X.; Kang, H. Downregulation of APE1 potentiates breast cancer cells to Olaparib by inhibiting PARP-1 expression. Breast Cancer Res. Treat. 2019, 176, 109–117. [Google Scholar] [CrossRef]
  47. Krutá, M.; Bálek, L.; Hejnová, R.; Dobšáková, Z.; Eiselleová, L.; Matulka, K.; Bárta, T.; Fojtík, P.; Fajkus, J.; Hampl, A.; et al. Decrease in abundance of apurinic/apyrimidinic endonuclease causes failure of base excision repair in culture-adapted human embryonic stem cells. Stem Cells 2013, 31, 693–702. [Google Scholar] [CrossRef]
  48. Heo, J.-Y.; Jing, K.; Song, K.-S.; Seo, K.-S.; Park, J.-H.; Kim, J.-S.; Jung, Y.-J.; Hur, G.-M.; Jo, D.-Y.; Kweon, G.-R.; et al. Downregulation of APE1/Ref-1 is involved in the senescence of mesenchymal stem cells. Stem Cells 2009, 27, 1455–1462. [Google Scholar] [CrossRef]
  49. Li, M.; Yang, X.; Lu, X.; Dai, N.; Zhang, S.; Cheng, Y.; Zhang, L.; Yang, Y.; Liu, Y.; Yang, Z.; et al. APE1 deficiency promotes cellular senescence and premature aging features. Nucleic Acids Res. 2018, 46, 5664–5677. [Google Scholar] [CrossRef]
  50. Xanthoudakis, S.; Smeyne, R.J.; Wallace, J.D.; Curran, T. The redox/DNA repair protein, Ref-1, is essential for early embryonic development in mice. Proc. Natl. Acad. Sci. USA 1996, 93, 8919–8923. [Google Scholar] [CrossRef] [Green Version]
  51. von Zglinicki, T. Oxidative stress shortens telomeres. Trends Biochem. Sci. 2002, 27, 339–344. [Google Scholar] [CrossRef]
  52. Meira, L.B.; Devaraj, S.; Kisby, G.E.; Burns, D.K.; Daniel, R.L.; Hammer, R.E.; Grundy, S.; Jialal, I.; Friedberg, E.C. Heterozygosity for the mouse apex gene results in phenotypes associated with oxidative stress. Cancer Res. 2001, 61, 5552–5557. [Google Scholar] [PubMed]
  53. Vogel, K.S.; Perez, M.; Momand, J.R.; Acevedo-Torres, K.; Hildreth, K.; Garcia, R.A.; Torres-Ramos, C.A.; Ayala-Torres, S.; Prihoda, T.J.; McMahan, C.A.; et al. Age-related instability in spermatogenic cell nuclear and mitochondrial DNA obtained fromApex1heterozygous mice. Mol. Reprod. Dev. 2011, 78, 906–919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Huamani, J.; McMahan, C.A.; Herbert, D.C.; Reddick, R.; McCarrey, J.R.; MacInnes, M.I.; Chen, D.J.; Walter, C.A. Spontaneous mutagenesis is enhanced in Apex heterozygous mice. Mol. Cell. Biol. 2004, 24, 8145–8153. [Google Scholar] [CrossRef] [Green Version]
  55. Li, M.; Wilson, D.M. Human apurinic/apyrimidinic endonuclease 1. Antioxid. Redox. Signal. 2014, 20, 678–707. [Google Scholar] [CrossRef] [Green Version]
  56. Abbotts, R.; Madhusudan, S. Human AP endonuclease 1 (APE1): From mechanistic insights to druggable target in cancer. Cancer Treat. Rev. 2010, 36, 425–435. [Google Scholar] [CrossRef]
  57. Karimi-Busheri, F.; Rasouli-Nia, A.; Mackey, J.R.; Weinfeld, M. Senescence evasion by MCF-7 human breast tumor-initiating cells. Breast Cancer Res. 2010, 12, R31. [Google Scholar] [CrossRef] [Green Version]
  58. Lin, Y.; Raj, J.; Li, J.; Ha, A.; Hossain, M.A.; Richardson, C.; Mukherjee, P.; Yan, S. APE1 senses DNA single-strand breaks for repair and signaling. Nucleic Acids Res. 2020, 48, 1925–1940. [Google Scholar] [CrossRef] [Green Version]
  59. Pieretti, M.; Khattar, N.H.; Smith, S.A. Common polymorphisms and somatic mutations in human base excision repair genes in ovarian and endometrial cancers. Mutat. Res. Res. Genom. 2001, 432, 53–59. [Google Scholar] [CrossRef]
  60. Illuzzi, J.L.; Harris, N.A.; Manvilla, B.A.; Kim, D.; Li, M.; Drohat, A.; Wilson, D.M. Functional assessment of population and tumor-associated APE1 protein variants. PLoS ONE 2013, 8, e65922. [Google Scholar] [CrossRef]
  61. Lo, Y.-L.; Jou, Y.-S.; Hsiao, C.-F.; Chang, G.-C.; Tsai, Y.-H.; Su, W.-C.; Chen, K.-Y.; Chen, Y.-M.; Huang, M.-S.; Hu, C.Y.; et al. A polymorphism in the APE1 gene promoter is associated with lung cancer risk. Cancer Epidemiol. Biomark. Prev. 2009, 18, 223–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lu, J.; Zhang, S.; Chen, D.; Wang, H.; Wu, W.; Wang, X.; Lei, Y.; Wang, J.; Qian, J.; Fan, W.; et al. Functional characterization of a promoter polymorphism in APE1/Ref-1 that contributes to reduced lung cancer susceptibility. FASEB J. 2009, 23, 3459–3469. [Google Scholar] [CrossRef]
  63. Ma, Y.; Pannicke, U.; Schwarz, K.; Lieber, M.R. Hairpin opening and overhang processing by an artemis/DNA-dependent protein kinase complex in nonhomologous end joining and V(D)J recombination. Cell 2002, 108, 781–794. [Google Scholar] [CrossRef] [Green Version]
  64. Li, S.; Chang, H.H.; Niewolik, D.; Hedrick, M.P.; Pinkerton, A.B.; Hassig, C.A.; Schwarz, K.; Lieber, M.R. Evidence that the DNA endonuclease ARTEMIS also has intrinsic 5′-exonuclease activity. J. Biol. Chem. 2014, 289, 7825–7834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Rooney, S.; Alt, F.W.; Lombard, D.; Whitlow, S.; Eckersdorff, M.; Fleming, J.; Fugmann, S.; Ferguson, D.O.; Schatz, D.G.; Sekiguchi, J. Defective DNA repair and increased genomic instability in artemis-deficient murine cells. J. Exp. Med. 2003, 197, 553–565. [Google Scholar] [CrossRef] [PubMed]
  66. Moshous, D.; Callebaut, I.; de Chasseval, R.; Corneo, B.; Cavazzana-Calvo, M.; Le Deist, F.; Tezcan, I.; Sanal, O.; Bertrand, Y.; Philippe, N.; et al. Artemis, a Novel DNA double-strand break repair/V(D)J recombination protein, is mutated in human severe combined immune deficiency. Cell 2001, 105, 177–186. [Google Scholar] [CrossRef] [Green Version]
  67. Kurosawa, A.; Koyama, H.; Takayama, S.; Miki, K.; Ayusawa, D.; Fujii, M.; Iiizumi, S.; Adachi, N. The requirement of artemis in double-strand break repair depends on the type of DNA damage. DNA Cell Biol. 2008, 27, 55–61. [Google Scholar] [CrossRef] [PubMed]
  68. Li, L.; Drayna, D.; Hu, D.; Hayward, A.; Gahagan, S.; Pabst, H.; Cowan, M.J. The gene for severe combined immunodeficiency disease in athabascan-speaking native americans is located on chromosome 10p. Am. J. Hum. Genet. 1998, 62, 136–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Murphy, S.; Trqup, G.; Hayward, A.; Devor, E.; Coons, T. Gene enrichment in an American Indian population: An excess of severe combined immunodeficiency disease. Lancet 1980, 316, 502–505. [Google Scholar] [CrossRef]
  70. Rooney, S.; Sekiguchi, J.; Zhu, C.; Cheng, H.-L.; Manis, J.; Whitlow, S.; DeVido, J.; Foy, D.; Chaudhuri, J.; Lombard, D.; et al. Leaky scid phenotype associated with defective V(D)J coding end processing in Artemis-deficient mice. Mol. Cell 2002, 10, 1379–1390. [Google Scholar] [CrossRef]
  71. Rooney, S.; Sekiguchi, J.; Whitlow, S.; Eckersdorff, M.; Manis, J.P.; Lee, C.; Ferguson, D.O.; Alt, F.W. Artemis and p53 cooperate to suppress oncogenic N-myc amplification in progenitor B cells. Proc. Natl. Acad. Sci. USA 2004, 101, 2410–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zhang, X.; Zhu, Y.; Geng, L.; Wang, H.; Legerski, R.J. Artemis is a negative regulator of p53 in response to oxidative stress. Oncogene 2009, 28, 2196–2204. [Google Scholar] [CrossRef] [Green Version]
  73. Sallmyr, A.; Tomkinson, A.E.; Rassool, F.V. Up-regulation of WRN and DNA ligase IIIalpha in chronic myeloid leukemia: Consequences for the repair of DNA double-strand breaks. Blood 2008, 112, 1413–1423. [Google Scholar] [CrossRef] [Green Version]
  74. Moshous, D.; Pannetier, C.; de Chasseval, R.; le Deist, F.; Cavazzana, M.; Romana, S.P.; Macintyre, E.; Canioni, D.; Brousse, N.; Fischer, A.; et al. Partial T and B lymphocyte immunodeficiency and predisposition to lymphoma in patients with hypomorphic mutations in Artemis. J. Clin. Investig. 2003, 111, 381–387. [Google Scholar] [CrossRef] [Green Version]
  75. Fevang, B.; Fagerli, U.M.; Sorte, H.; Aarset, H.; Hov, H.; Langmyr, M.; Keil, T.M.; Bjørge, E.; Aukrust, P.; Stray-Pedersen, A.; et al. Runaway train: A leaky radiosensitive SCID with skin lesions and multiple lymphomas. Case Rep. Immunol. 2018, 2053716. [Google Scholar] [CrossRef] [PubMed]
  76. Jacobs, C.; Huang, Y.; Masud, T.; Lu, W.; Westfield, G.; Giblin, W.; Sekiguchi, J.M. A hypomorphic Artemis human disease allele causes aberrant chromosomal rearrangements and tumorigenesis. Hum. Mol. Genet. 2010, 20, 806–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhu, C.; Wang, X.; Li, P.; Zhu, Y.; Sun, Y.; Hu, J.; Liu, H.; Sun, X. Developing a peptide that inhibits DNA repair by blocking the binding of Artemis and DNA ligase IV to enhance tumor radiosensitivity. Int. J. Radiat. Oncol. Biol. Phys. 2021, 111, 515–527. [Google Scholar] [CrossRef]
  78. Park, J.; Lee, S.-Y.; Jeong, H.; Kang, M.-G.; Van Haute, L.; Minczuk, M.; Seo, J.K.; Jun, Y.; Myung, K.; Rhee, H.-W.; et al. The structure of human EXD2 reveals a chimeric 3′ to 5′ exonuclease domain that discriminates substrates via metal coordination. Nucleic Acids Res. 2019, 47, 7078–7093. [Google Scholar] [CrossRef] [Green Version]
  79. Zid, B.M.; Kapahi, P. Exonuclease EXD2 in mitochondrial translation. Nat. Cell Biol. 2018, 20, 120–122. [Google Scholar] [CrossRef]
  80. Nieminuszczy, J.; Broderick, R.; Bellani, M.A.; Smethurst, E.; Schwab, R.A.; Cherdyntseva, V.; Evmorfopoulou, T.; Lin, Y.-L.; Minczuk, M.; Pasero, P.; et al. EXD2 protects stressed replication forks and is required for cell viability in the absence of BRCA1/2. Mol. Cell 2019, 75, 605–619.e6. [Google Scholar] [CrossRef] [Green Version]
  81. Broderick, R.; Nieminuszczy, J.; Baddock, H.T.; Deshpande, R.A.; Gileadi, O.; Paull, T.T.; McHugh, P.J.; Niedzwiedz, W. EXD2 promotes homologous recombination by facilitating DNA end resection. Nat. Cell Biol. 2016, 18, 271–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Sertic, S.; Quadri, R.; Lazzaro, F.; Muzi-Falconi, M. EXO1: A tightly regulated nuclease. DNA Repair 2020, 93, 102929. [Google Scholar] [CrossRef] [PubMed]
  83. Goellner, E.M.; Putnam, C.D.; Kolodner, R.D. Exonuclease 1-dependent and independent mismatch repair. DNA Repair 2015, 32, 24–32. [Google Scholar] [CrossRef] [Green Version]
  84. Bolderson, E.; Tomimatsu, N.; Richard, D.J.; Boucher, D.; Kumar, R.; Pandita, T.K.; Burma, S.; Khanna, K.K. Phosphorylation of Exo1 modulates homologous recombination repair of DNA double-strand breaks. Nucleic Acids Res. 2010, 38, 1821–1831. [Google Scholar] [CrossRef] [Green Version]
  85. Schaetzlein, S.; Chahwan, R.; Avdievich, E.; Roa, S.; Wei, K.; Eoff, R.L.; Sellers, R.S.; Clark, A.B.; Kunkel, T.A.; Scharff, M.D.; et al. Mammalian Exo1 encodes both structural and catalytic functions that play distinct roles in essential biological processes. Proc. Natl. Acad. Sci. USA 2013, 110, E2470–E2479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Mimitou, E.P.; Symington, L.S. Sae2, Exo1 and Sgs1 collaborate in DNA double-strand break processing. Nature 2008, 455, 770–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Sertic, S.; Mollica, A.; Campus, I.; Roma, S.; Tumini, E.; Aguilera, A.; Muzi-Falconi, M. Coordinated activity of Y family TLS polymerases and EXO1 protects Non-S phase cells from UV-induced cytotoxic lesions. Mol. Cell 2018, 70, 34–47.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Giannattasio, M.; Follonier, C.; Tourrière, H.; Puddu, F.; Lazzaro, F.; Pasero, P.; Lopes, M.; Plevani, P.; Muzi-Falconi, M. Exo1 competes with repair synthesis, converts Ner intermediates to long ssDNA gaps, and promotes checkpoint activation. Mol. Cell 2010, 40, 50–62. [Google Scholar] [CrossRef] [PubMed]
  89. Luo, F.; Wang, Y.; Lin, D.; Li, J.; Yang, K. Exonuclease 1 expression is associated with clinical progression, metastasis, and survival prognosis of prostate cancer. J. Cell. Biochem. 2019, 120, 11383–11389. [Google Scholar] [CrossRef]
  90. Teng, P.-C.; Huang, S.-P.; Liu, C.-H.; Lin, T.-Y.; Cho, Y.-C.; Lai, Y.-L.; Wang, S.-C.; Yeh, H.-C.; Chuu, C.-P.; Chen, D.-N.; et al. Identification of DNA damage repair-associated prognostic biomarkers for prostate cancer using transcriptomic data analysis. Int. J. Mol. Sci. 2021, 22, 11771. [Google Scholar] [CrossRef] [PubMed]
  91. Liu, J.; Zhang, J. Elevated EXO1 expression is associated with breast carcinogenesis and poor prognosis. Ann. Transl. Med. 2021, 9, 135. [Google Scholar] [CrossRef] [PubMed]
  92. Li, Y.; Wang, Y.; Zhang, W.; Wang, X.; Chen, L.; Wang, S. BKM120 sensitizes BRCA-proficient triple negative breast cancer cells to olaparib through regulating FOXM1 and Exo1 expression. Sci. Rep. 2021, 11, 4774. [Google Scholar] [CrossRef] [PubMed]
  93. Huang, K.; Wu, Y.; Xie, Y.; Huang, L.; Liu, H. Analyzing mRNAsi-related genes identifies novel prognostic markers and potential drug combination for patients with basal breast cancer. Dis. Markers 2021, 2021, 4731349. [Google Scholar] [CrossRef]
  94. He, D.; Li, T.; Sheng, M.; Yang, B. Exonuclease 1 (Exo1) participates in mammalian non-homologous end joining and contributes to drug resistance in ovarian cancer. Med Sci. Monit. 2020, 26, e918751-1–e918751-8. [Google Scholar] [CrossRef] [PubMed]
  95. Zhou, C.-S.; Feng, M.-T.; Chen, X.; Gao, Y.; Chen, L.; Li, L.-D.; Li, D.-H.; Cao, Y.-Q. Exonuclease 1 (EXO1) is a potential prognostic biomarker and correlates with immune infiltrates in lung adenocarcinoma. OncoTargets Ther. 2021, 14, 1033–1048. [Google Scholar] [CrossRef] [PubMed]
  96. Dai, Y.; Tang, Z.; Yang, Z.; Zhang, L.; Deng, Q.; Zhang, X.; Yu, Y.; Liu, X.; Zhu, J. EXO1 overexpression is associated with poor prognosis of hepatocellular carcinoma patients. Cell Cycle 2018, 17, 2386–2397. [Google Scholar] [CrossRef]
  97. Yang, G.; Dong, K.; Zhang, Z.; Zhang, E.; Liang, B.; Chen, X.; Huang, Z. EXO1 plays a carcinogenic role in hepatocellular carcinoma and is related to the regulation of FOXP3. J. Cancer 2020, 11, 4917–4932. [Google Scholar] [CrossRef]
  98. Fan, J.; Zhao, Y.; Yuan, H.; Yang, J.; Li, T.; He, Z.; Wu, X.; Luo, C. Phospholipase C-ε regulates bladder cancer cells via ATM/EXO1. Am. J. Cancer Res. 2020, 10, 2319–2336. [Google Scholar]
  99. Song, F.; Qureshi, A.A.; Zhan, J.; Amos, C.I.; Lee, J.E.; Wei, Q.; Han, J. Exonuclease 1 (EXO1) gene variation and melanoma risk. DNA Repair 2012, 11, 304–309. [Google Scholar] [CrossRef] [Green Version]
  100. Sun, X.; Zheng, L.; Shen, B. Functional alterations of human exonuclease 1 mutants identified in atypical hereditary nonpolyposis colorectal cancer syndrome. Cancer Res. 2002, 62, 6026–6030. [Google Scholar]
  101. Zhang, Y.; Li, P.; Xu, A.; Chen, J.; Ma, C.; Sakai, A.; Xie, L.; Wang, L.; Na, Y.; Kaku, H.; et al. Influence of a single-nucleotide polymorphism of the DNA mismatch repair-related gene exonuclease-1 (rs9350) with prostate cancer risk among Chinese people. Tumor Biol. 2015, 37, 6653–6659. [Google Scholar] [CrossRef]
  102. Shi, T.; Jiang, R.; Wang, P.; Xu, Y.; Yin, S.; Cheng, X.; Zang, R. Significant association of the EXO1 rs851797 polymorphism with clinical outcome of ovarian cancer. OncoTargets Ther. 2017, 10, 4841–4851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Jin, G.; Wang, H.; Hu, Z.; Liu, H.; Sun, W.; Ma, H.; Chen, D.; Miao, R.; Tian, T.; Jin, L.; et al. Potentially functional polymorphisms of EXO1 and risk of lung cancer in a Chinese population: A case-control analysis. Lung Cancer 2008, 60, 340–346. [Google Scholar] [CrossRef] [PubMed]
  104. Duan, F.; Song, C.; Dai, L.; Cui, S.; Zhang, X.; Zhao, X. The significance of Exo1 K589E polymorphSism on cancer susceptibility: Evidence based on a meta-analysis. PLoS ONE 2014, 9, e96764. [Google Scholar] [CrossRef]
  105. Genetic Risk of Lung Cancer Associated with a Single Nucleotide Polymorphism from EXO1: A Meta Analysis—PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/26379914/ (accessed on 15 April 2022).
  106. Tsai, M.-H.; Tseng, H.-C.; Liu, C.-S.; Chang, C.-L.; Tsai, C.-W.; Tsou, Y.-A.; Wang, R.-F.; Lin, C.-C.; Wang, H.-C.; Chiu, C.-F.; et al. Interaction of Exo1 genotypes and smoking habit in oral cancer in Taiwan. Oral Oncol. 2009, 45, e90–e94. [Google Scholar] [CrossRef]
  107. Bayram, S.; Akkiz, H.; Bekar, A.; Akgöllü, E.; Yildirim, S. The significance of Exonuclease 1 K589E polymorphism on hepatocellular carcinoma susceptibility in the Turkish population: A case-control study. Mol. Biol. Rep.Mol. Biol. Rep. 2012, 39, 5943–5951. [Google Scholar] [CrossRef]
  108. Wu, Y.; Mensink, R.G.; Verlind, E.; Sijmons, R.; Buys, C.H.; Hofstra, R.M.; Berends, M.J.; Kleibeuker, J.H.; Post, J.G.; Kempinga, C.; et al. Germline mutations of EXO1 gene in patients with hereditary nonpolyposis colorectal cancer (HNPCC) and atypical HNPCC forms. Gastroenterology 2001, 120, 1580–1587. [Google Scholar] [CrossRef]
  109. Bau, D.-T. Single-nucleotide polymorphism of the Exo1 gene: Association with gastric cancer susceptibility and interaction with smoking in Taiwan. Chin. J. Physiol. 2009, 52, 411–418. [Google Scholar] [CrossRef] [PubMed]
  110. Tan, S.; Qin, R.; Zhu, X.; Tan, C.; Song, J.; Qin, L.; Liu, L.; Huang, X.; Li, A.; Qiu, X. Associations between single-nucleotide polymorphisms of human exonuclease 1 and the risk of hepatocellular carcinoma. Oncotarget 2016, 7, 87180–87193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Haghighi, M.M.; Taleghani, M.Y.; Mohebbi, S.R.; Vahedi, M.; Fatemi, S.R.; Zali, N.; Shemirani, A.I.; Zali, M.R. Impact of EXO1 polymorphism in susceptibility to colorectal cancer. Genet. Test. Mol. Biomark. 2010, 14, 649–652. [Google Scholar] [CrossRef]
  112. Wei, K.; Clark, A.B.; Wong, E.; Kane, M.F.; Mazur, D.J.; Parris, T.; Kolas, N.K.; Russell, R.; Hou, H., Jr.; Kneitz, B.; et al. Inactivation of exonuclease 1 in mice results in DNA mismatch repair defects, increased cancer susceptibility, and male and female sterility. Genes Dev. 2003, 17, 603–614. [Google Scholar] [CrossRef] [Green Version]
  113. Guan, J.; Lu, C.; Jin, Q.; Lu, H.; Chen, X.; Tian, L.; Zhang, Y.; Rodriguez, J.O.; Zhang, J.; Siteni, S.; et al. MLH1 deficiency-triggered DNA Hyperexcision by exonuclease 1 activates the cGAS-STING pathway. Cancer Cell 2021, 39, 109–121.e5. [Google Scholar] [CrossRef] [PubMed]
  114. Kieper, J.; Lauber, C.; Gimadutdinow, O.; Urbańska, A.; Cymerman, I.; Ghosh, M.; Szczesny, B.; Meiss, G. Production and characterization of recombinant protein preparations of endonuclease G-homologs from yeast, C. elegans and humans. Protein Expr. Purif. 2010, 73, 99–106. [Google Scholar] [CrossRef] [PubMed]
  115. Cymerman, I.A.; Chung, I.; Beckmann, B.M.; Bujnicki, J.M.; Meiss, G. EXOG, a novel paralog of Endonuclease G in higher eukaryotes. Nucleic Acids Res. 2008, 36, 1369–1379. [Google Scholar] [CrossRef] [Green Version]
  116. Tann, A.W.; Boldogh, I.; Meiss, G.; Qian, W.; Van Houten, B.; Mitra, S.; Szczesny, B. Apoptosis induced by persistent single-strand breaks in mitochondrial genome: Critical role of EXOG (5′-EXO/endonuclease) in their repair. J. Biol. Chem. 2011, 286, 31975–31983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Szczesny, B.; Marcatti, M.; Zatarain, J.R.; Druzhyna, N.; Wiktorowicz, J.E.; Nagy, P.; Hellmich, M.R.; Szabo, C. Inhibition of hydrogen sulfide biosynthesis sensitizes lung adenocarcinoma to chemotherapeutic drugs by inhibiting mitochondrial DNA repair and suppressing cellular bioenergetics. Sci. Rep. 2016, 6, 36125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Wu, C.-C.; Lin, J.L.J.; Yang-Yen, H.-F.; Yuan, H.S. A unique exonuclease ExoG cleaves between RNA and DNA in mitochondrial DNA replication. Nucleic Acids Res. 2019, 47, 5405–5419. [Google Scholar] [CrossRef] [Green Version]
  119. Lung, M.S.; Australian Ovarian Cancer Study Group; Mitchell, C.A.; Doyle, M.A.; Lynch, A.C.; Gorringe, K.L.; Bowtell, D.D.L.; Campbell, I.G.; Trainer, A.H. Germline whole exome sequencing of a family with appendiceal mucinous tumours presenting with pseudomyxoma peritonei. BMC Cancer 2020, 20, 369. [Google Scholar] [CrossRef]
  120. Jin, H.; Cho, Y. Structural and functional relationships of FAN1. DNA Repair 2017, 56, 135–143. [Google Scholar] [CrossRef]
  121. Fiévet, A.; Mouret-Fourme, E.; Colas, C.; de Pauw, A.; Stoppa-Lyonnet, D.; Buecher, B. Prevalence of pathogenic variants of FAN1 in more than 5000 patients assessed for genetic predisposition to colorectal, breast, ovarian, or other cancers. Gastroenterology 2019, 156, 1919–1920. [Google Scholar] [CrossRef]
  122. Smith, A.L.; Alirezaie, N.; Connor, A.; Chan-Seng-Yue, M.; Grant, R.; Selander, I.; Bascuñana, C.; Borgida, A.; Hall, A.; Whelan, T.; et al. Candidate DNA repair susceptibility genes identified by exome sequencing in high-risk pancreatic cancer. Cancer Lett. 2015, 370, 302–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Belhadj, S.; Terradas, M.; Munoz-Torres, P.M.; Aiza, G.; Navarro, M.; Capellá, G.; Valle, L. Candidate genes for hereditary colorectal cancer: Mutational screening and systematic review. Hum. Mutat. 2020, 41, 1563–1576. [Google Scholar] [CrossRef] [PubMed]
  124. Seguí, N.; Mina, L.B.; Lázaro, C.; Sanz-Pamplona, R.; Pons, T.; Navarro, M.; Bellido, F.; López-Doriga, A.; Valdés-Mas, R.; Pineda, M.; et al. Germline mutations in FAN1 cause hereditary colorectal cancer by impairing DNA repair. Gastroenterology 2015, 149, 563–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hou, P.; Su, X.; Cao, W.; Xu, L.; Zhang, R.; Huang, Z.; Wang, J.; Li, L.; Wu, L.; Liao, W. Whole-exome sequencing reveals the etiology of the rare primary hepatic mucoepidermoid carcinoma. Diagn. Pathol. 2021, 16, 29. [Google Scholar] [CrossRef]
  126. Lachaud, C.; Moreno, A.; Marchesi, F.; Toth, R.; Blow, J.J.; Rouse, J. Ubiquitinated Fancd2 recruits Fan1 to stalled replication forks to prevent genome instability. Science 2016, 351, 846–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Porro, A.; Berti, M.; Pizzolato, J.; Bologna, S.; Kaden, S.; Saxer, A.; Ma, Y.; Nagasawa, K.; Sartori, A.A.; Jiricny, J. FAN1 interaction with ubiquitylated PCNA alleviates replication stress and preserves genomic integrity independently of BRCA2. Nat. Commun. 2017, 8, 1073. [Google Scholar] [CrossRef] [Green Version]
  128. Santarpia, L.; Iwamoto, T.; Di Leo, A.; Hayashi, N.; Bottai, G.; Stampfer, M.; André, F.; Turner, N.C.; Symmans, W.F.; Hortobágyi, G.N.; et al. DNA repair gene patterns as prognostic and predictive factors in molecular breast cancer subtypes. Oncology 2013, 18, 1063–1073. [Google Scholar] [CrossRef] [Green Version]
  129. Zheng, L.; Dai, H.; Zhou, M.; Li, M.; Singh, P.; Qiu, J.; Tsark, W.; Huang, Q.; Kernstine, K.; Zhang, X.; et al. Fen1 mutations result in autoimmunity, chronic inflammation and cancers. Nat. Med. 2007, 13, 812–819. [Google Scholar] [CrossRef]
  130. Larsen, E.; Gran, C.; Sæther, B.E.; Seeberg, E.; Klungland, A. Proliferation failure and gamma radiation sensitivity of Fen1 null mutant mice at the blastocyst stage. Mol. Cell. Biol. 2003, 23, 5346–5353. [Google Scholar] [CrossRef] [Green Version]
  131. Kucherlapati, M.; Yang, K.; Kuraguchi, M.; Zhao, J.; Lia, M.; Heyer, J.; Kane, M.F.; Fan, K.; Russell, R.; Brown, A.M.C.; et al. Haploinsufficiency of Flap endonuclease (Fen1) leads to rapid tumor progression. Proc. Natl. Acad. Sci. USA 2002, 99, 9924–9929. [Google Scholar] [CrossRef] [Green Version]
  132. Sun, H.; He, L.; Wu, H.; Pan, F.; Wu, X.; Zhao, J.; Hu, Z.; Sekhar, C.; Li, H.; Zheng, L.; et al. The FEN1 L209P mutation interferes with long-patch base excision repair and induces cellular transformation. Oncogene 2016, 36, 194–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Lam, J.S.; Seligson, D.B.; Yu, H.; Li, A.; Eeva, M.; Pantuck, A.J.; Zeng, G.; Horvath, S.; Belldegrun, A.S. Flap endonuclease 1 is overexpressed in prostate cancer and is associated with a high Gleason score. Br. J. Urol. 2006, 98, 445–451. [Google Scholar] [CrossRef] [PubMed]
  134. Nikolova, T.; Christmann, M.; Kaina, B. FEN1 is overexpressed in testis, lung and brain tumors. Anticancer Res. 2009, 29, 2453–2459. [Google Scholar] [PubMed]
  135. Sato, M.; Girard, L.; Sekine, I.; Sunaga, N.; Ramirez, R.D.; Kamibayashi, C.; Minna, J.D. Increased expression and no mutation of the Flap endonuclease (FEN1) gene in human lung cancer. Oncogene 2003, 22, 7243–7246. [Google Scholar] [CrossRef] [Green Version]
  136. Wang, K.; Xie, C.; Chen, D. Flap endonuclease 1 is a promising candidate biomarker in gastric cancer and is involved in cell proliferation and apoptosis. Int. J. Mol. Med. 2014, 33, 1268–1274. [Google Scholar] [CrossRef] [Green Version]
  137. Singh, P.; Yang, M.; Dai, H.; Yu, D.; Huang, Q.; Tan, W.; Kernstine, K.H.; Lin, D.; Shen, B. Overexpression and hypomethylation of Flap Endonuclease 1 gene in breast and other cancers. Mol. Cancer Res. 2008, 6, 1710–1717. [Google Scholar] [CrossRef]
  138. Abdel-Fatah, T.M.; Russell, R.; Albarakati, N.; Maloney, D.J.; Dorjsuren, D.; Rueda, O.M.; Moseley, P.; Mohan, V.; Sun, H.; Abbotts, R.; et al. Genomic and protein expression analysis reveals flap endonuclease 1 (FEN1) as a key biomarker in breast and ovarian cancer. Mol. Oncol. 2014, 8, 1326–1338. [Google Scholar] [CrossRef]
  139. Liu, L.; Zhou, C.; Zhou, L.; Peng, L.; Li, D.; Zhang, X.; Zhou, M.; Kuang, P.; Yuan, Q.; Song, X.; et al. Functional FEN1 genetic variants contribute to risk of hepatocellular carcinoma, esophageal cancer, gastric cancer and colorectal cancer. Carcinogenesis 2011, 33, 119–123. [Google Scholar] [CrossRef] [Green Version]
  140. Chen, Y.-D.; Zhang, X.; Qiu, X.-G.; Li, J.; Yuan, Q.; Jiang, T.; Yang, M. Functional FEN1 genetic variants and haplotypes are associated with glioma risk. J. Neuro-Oncol. 2012, 111, 145–151. [Google Scholar] [CrossRef]
  141. Lv, Z.; Liu, W.; Li, D.; Liu, L.; Wei, J.; Zhang, J.; Ge, Y.; Wang, Z.; Chen, H.; Zhou, C.; et al. Association of functional FEN1 genetic variants and haplotypes and breast cancer risk. Gene 2014, 538, 42–45. [Google Scholar] [CrossRef]
  142. Ren, H.; Ma, H.; Ke, Y.; Ma, X.; Xu, D.; Lin, S.; Wang, X.; Dai, Z.-J. Flap endonuclease 1 polymorphisms (rs174538 and rs4246215) contribute to an increased cancer risk: Evidence from a meta-analysis. Mol. Clin. Oncol. 2015, 3, 1347–1352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Jiao, X.; Wu, Y.; Zhou, L.; He, J.; Yang, C.; Zhang, P.; Hu, R.; Luo, C.; Du, J.; Fu, J.; et al. Variants and haplotypes in Flap endonuclease 1 and risk of gallbladder cancer and gallstones: A population-based study in China. Sci. Rep. 2015, 5, 18160. [Google Scholar] [CrossRef] [Green Version]
  144. Pei, J.S.; Chang, W.S.; Hsu, P.C.; Tsai, C.W.; Hsu, C.M.; Ji, H.X.; Hsiao, C.L.; Hsu, Y.N.; Bau, D.T. The association of flap endonuclease 1 genotypes with the risk of childhood leukemia. Cancer Genom. Proteomics. 2016, 13, 69–74. [Google Scholar]
  145. Rezaei, M.; Hashemi, M.; Sanaei, S.; Mashhadi, M.A.; Hashemi, S.M.; Bahari, G.; Taheri, M. FEN1 −69G>A and +4150G>T polymorphisms and breast cancer risk. Biomed. Rep. 2016, 5, 455–460. [Google Scholar] [CrossRef] [Green Version]
  146. Lin, S.; Wang, M.; Liu, X.; Lu, Y.; Gong, Z.; Guo, Y.; Yang, P.; Tian, T.; Dai, C.; Zheng, Y.; et al. FEN1 gene variants confer reduced risk of breast cancer in Chinese women: A case-control study. Oncotarget 2016, 7, 78110–78118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Sang, Y.; Bo, L.; Gu, H.; Yang, W.; Chen, Y. Flap endonuclease-1 rs174538 G>A polymorphisms are associated with the risk of esophageal cancer in a Chinese population. Thorac. Cancer 2017, 8, 192–196. [Google Scholar] [CrossRef] [PubMed]
  148. Krupa, R.; Czarny, P.; Wigner, P.; Wozny, J.; Jablkowski, M.; Kordek, R.; Szemraj, J.; Sliwinski, T. The relationship between single-nucleotide polymorphisms, the expression of DNA damage response genes, and hepatocellular carcinoma in a Polish population. DNA Cell Biol. 2017, 36, 693–708. [Google Scholar] [CrossRef]
  149. Moazeni-Roodi, A.; Ghavami, S.; Ansari, H.; Hashemi, M. Association between the flap endonuclease 1 gene polymorphisms and cancer susceptibility: An updated meta-analysis. J. Cell. Biochem. 2019, 120, 13583–13597. [Google Scholar] [CrossRef]
  150. Zhang, M.; Zhao, Z.; Chen, S.; Liang, Z.; Zhu, J.; Zhao, M.; Xu, C.; He, J.; Duan, P.; Zhang, A. The association of polymorphisms in base excision repair genes with ovarian cancer susceptibility in Chinese women: A two-center case-control study. J. Cancer 2021, 12, 264–269. [Google Scholar] [CrossRef]
  151. Yang, M.; Guo, H.; Wu, C.; He, Y.; Yu, D.; Zhou, L.; Wang, F.; Xu, J.; Tan, W.; Wang, G.; et al. FunctionalFEN1polymorphisms are associated with DNA damage levels and lung cancer risk. Hum. Mutat. 2009, 30, 1320–1328. [Google Scholar] [CrossRef]
  152. Ying, N.; Wang, S.; Xu, H.; Wang, Y. Association between FEN1 Polymorphisms -69G>A and 4150G>T with Susceptibility in Human Disease: A Meta-Analysis. Iran. J. Public Health 2015, 44, 1574–1579. [Google Scholar] [PubMed]
  153. Reginato, G.; Cejka, P. The MRE11 complex: A versatile toolkit for the repair of broken DNA. DNA Repair 2020, 91–92, 102869. [Google Scholar] [CrossRef] [PubMed]
  154. Damiola, F.; Pertesi, M.; Oliver, J.; Le Calvez-Kelm, F.; Voegele, C.; Young, E.L.; Robinot, N.; Forey, N.; Durand, G.; Vallée, M.P.; et al. Rare key functional domain missense substitutions in MRE11A, RAD50, and NBN contribute to breast cancer susceptibility: Results from a breast cancer family registry case-control mutation-screening study. Breast Cancer Res. 2014, 16, R58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Rahman, S.; Beikzadeh, M.; Latham, M.P. Biochemical and structural characterization of analogs of MRE11 breast cancer-associated mutant F237C. Sci. Rep. 2021, 11, 7089. [Google Scholar] [CrossRef] [PubMed]
  156. Wen, Q.; Scorah, J.; Phear, G.; Rodgers, G.; Rodgers, S.; Meuth, M. A Mutant allele of MRE11 found in mismatch repair-deficient tumor cells suppresses the cellular response to DNA replication fork stress in a dominant negative manner. Mol. Biol. Cell 2008, 19, 1693–1705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Giannini, G.; Rinaldi, C.; Ristori, E.; Ambrosini, M.I.; Cerignoli, F.; Viel, A.; Bidoli, E.; Berni, S.; D’Amati, G.; Scambia, G.; et al. Mutations of an intronic repeat induce impaired MRE11 expression in primary human cancer with microsatellite instability. Oncogene 2004, 23, 2640–2647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Rahman, S.; Beikzadeh, M.; Canny, M.D.; Kaur, N.; Latham, M.P. Mutation of conserved mre11 residues alter protein dynamics to separate nuclease functions. J. Mol. Biol. 2020, 432, 3289–3308. [Google Scholar] [CrossRef]
  159. Choudhury, A.; Nelson, L.D.; Teo, M.T.; Chilka, S.; Bhattarai, S.; Johnston, C.F.; Elliott, F.; Lowery, J.; Taylor, C.F.; Churchman, M.; et al. MRE11 Expression is predictive of cause-specific survival following radical radiotherapy for muscle-invasive bladder cancer. Cancer Res. 2010, 70, 7017–7026. [Google Scholar] [CrossRef] [Green Version]
  160. Mummenbrauer, T.; Janus, F.; Müller, B.; Wiesmüller, L.; Deppert, W.; Grosse, F. p53 Protein exhibits 3′-to-5′ exonuclease activity. Cell 1996, 85, 1089–1099. [Google Scholar] [CrossRef] [Green Version]
  161. Ahn, J.; Poyurovsky, M.V.; Baptiste, N.; Beckerman, R.; Cain, C.; Mattia, M.; McKinney, K.; Zhou, J.; Zupnick, A.; Gottifredi, V.; et al. Dissection of the sequence-specific DNA binding and exonuclease activities reveals a superactive yet apoptotically impaired mutant p53 protein. Cell Cycle 2009, 8, 1603–1615. [Google Scholar] [CrossRef]
  162. Ho, T.; Tan, B.X.; Lane, D. How the other half lives: What p53 does when it is not being a transcription factor. Int. J. Mol. Sci. 2019, 21, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Chumakov, P.M. Versatile functions of p53 protein in multicellular organisms. Biochemistry 2007, 72, 1399–1421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Janus, F.; Albrechtsen, N.; Knippschild, U.; Wiesmüller, L.; Grosse, F.; Deppert, W. Different regulation of the p53 core domain activities 3′-to-5′ exonuclease and sequence-specific DNA binding. Mol. Cell. Biol. 1999, 19, 2155–2168. [Google Scholar] [CrossRef] [Green Version]
  165. Gila, L.; Elena, N.; Yechezkel, S.; Mary, B. p53-associated 3′–>5′ exonuclease activity in nuclear and cytoplasmic compartments of cells. Oncogene 2003, 22, 233–245. [Google Scholar] [CrossRef] [Green Version]
  166. Janus, F.; Albrechtsen, N.; Dornreiter, I.; Wiesmuller, L.; Grosse, F.; Deppert, W. The dual role model for p53 in maintaining genomic integrity. Cell Mol. Life Sci. 1999, 55, 12–27. [Google Scholar] [CrossRef] [PubMed]
  167. Schmale, H.; Bamberger, C. A novel protein with strong homology to the tumor suppressor p53. Oncogene 1997, 15, 1363–1367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Bakhanashvili, M.; Gedelovich, R.; Grinberg, S.; Rahav, G. Exonucleolytic degradation of RNA by p53 protein in cytoplasm. Klin. Wochenschr. 2007, 86, 75–88. [Google Scholar] [CrossRef]
  169. Grinberg, S.; Teiblum, G.; Rahav, G.; Bakhanashvili, M. p53 in cytoplasm exerts 3′→5′ exonuclease activity with dsRNA. Cell Cycle 2010, 9, 2442–2455. [Google Scholar] [CrossRef]
  170. Jenkins, G.M.; Frohman, M.A. Phospholipase D: A lipid centric review. Cell Mol. Life Sci. 2005, 62, 2305–2316. [Google Scholar] [CrossRef]
  171. Brown, H.A.; Thomas, P.G.; Lindsley, C.W. Targeting phospholipase D in cancer, infection and neurodegenerative disorders. Nat. Rev. Drug Discov. 2017, 16, 351–367. [Google Scholar] [CrossRef]
  172. Akizuki, S.; Ishigaki, K.; Kochi, Y.; Law, S.-M.; Matsuo, K.; Ohmura, K.; Suzuki, A.; Nakayama, M.; Iizuka, Y.; Koseki, H.; et al. PLD4 is a genetic determinant to systemic lupus erythematosus and involved in murine autoimmune phenotypes. Ann. Rheum. Dis. 2019, 78, 509–518. [Google Scholar] [CrossRef] [PubMed]
  173. Cruchaga, C.; Karch, C.M.; Jin, S.C.; Benitez, B.A.; Cai, Y.; Guerreiro, R.; Harari, O.; Norton, J.; Budde, J.; Bertelsen, S.; et al. Rare coding variants in the phospholipase D3 gene confer risk for Alzheimer’s disease. Nature 2014, 505, 550–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Guo, J.; Li, X.; Shen, S.; Wu, X. Expression of immune-related genes as prognostic biomarkers for the assessment of osteosarcoma clinical outcomes. Sci. Rep. 2021, 11, 24123. [Google Scholar] [CrossRef] [PubMed]
  175. Fan, L.; Ru, J.; Liu, T.; Ma, C. Identification of a novel prognostic gene signature from the immune cell infiltration landscape of osteosarcoma. Front. Cell Dev. Biol. 2021, 9, 2394. [Google Scholar] [CrossRef]
  176. Zhou, D.; Wu, Y.; Jiang, K.; Xu, F.; Hong, R.; Wang, S. Identification of a risk prediction model for clinical prognosis in HER2 positive breast cancer patients. Genomics 2021, 113, 4088–4097. [Google Scholar] [CrossRef]
  177. Gao, L.; Zhou, Y.; Zhou, S.-X.; Yu, X.-J.; Xu, J.-M.; Zuo, L.; Luo, Y.-H.; Li, X.-A. PLD4 promotes M1 macrophages to perform antitumor effects in colon cancer cells. Oncol. Rep. 2016, 37, 408–416. [Google Scholar] [CrossRef] [Green Version]
  178. Goldsby, R.E.; Hays, L.E.; Chen, X.; Olmsted, E.A.; Slayton, W.B.; Spangrude, G.J.; Preston, B.D. High incidence of epithelial cancers in mice deficient for DNA polymerase delta proofreading. Proc. Natl. Acad. Sci. USA 2002, 99, 15560–15565. [Google Scholar] [CrossRef] [Green Version]
  179. Park, V.S.; Pursell, Z.F. POLE proofreading defects: Contributions to mutagenesis and cancer. DNA Repair 2019, 76, 50–59. [Google Scholar] [CrossRef]
  180. Magrin, L.; Fanale, D.; Brando, C.; Fiorino, A.; Corsini, L.R.; Sciacchitano, R.; Filorizzo, C.; Dimino, A.; Russo, A.; Bazan, V. POLE, POLD1, and NTHL1: The last but not the least hereditary cancer-predisposing genes. Oncogene 2021, 40, 5893–5901. [Google Scholar] [CrossRef]
  181. Mur, P.; Ms, S.G.-M.; del Valle, J.; Ms, L.M.-P.; Vidal, A.; Pineda, M.; Cinnirella, G.; Ms, E.M.-R.; Pons, T.; López-Doriga, A.; et al. Role of POLE and POLD1 in familial cancer. Genet. Med. 2020, 22, 2089–2100. [Google Scholar] [CrossRef]
  182. Broustas, C.G.; Lieberman, H.B. Contributions of Rad9 to tumorigenesis. J. Cell. Biochem. 2011, 113, 742–751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Roos-Mattjus, P.; Hopkins, K.M.; Oestreich, A.J.; Vroman, B.T.; Johnson, K.L.; Naylor, S.; Lieberman, H.B.; Karnitz, L.M. Phosphorylation of human Rad9 is required for genotoxin-activated checkpoint signaling. J. Biol. Chem. 2003, 278, 24428–24437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Maniwa, Y.; Yoshimura, M.; Bermudez, V.P.; Yuki, T.; Okada, K.; Kanomata, N.; Ohbayashi, C.; Hayashi, Y.; Hurwitz, J.; Okita, Y. Accumulation of hRad9 protein in the nuclei of nonsmall cell lung carcinoma cells. Cancer 2004, 103, 126–132. [Google Scholar] [CrossRef] [PubMed]
  185. Maniwa, Y.; Yuki, T.; Doi, T.; Okada, K.; Nishio, W.; Hayashi, Y.; Okita, Y. DNA damage sensor protein hRad9, a novel molecular target for lung cancer treatment. Oncol. Rep. 1994, 20, 1047–1052. [Google Scholar] [CrossRef]
  186. Kebebew, E.; Peng, M.; Reiff, E.; Duh, Q.-Y.; Clark, O.H.; McMillan, A. Diagnostic and prognostic value of cell-cycle regulatory genes in malignant thyroid neoplasms. World J. Surg. 2006, 30, 767–774. [Google Scholar] [CrossRef]
  187. Zhu, A.; Zhang, C.X.; Lieberman, H.B. Rad9 has a functional role in human prostate carcinogenesis. Cancer Res. 2008, 68, 1267–1274. [Google Scholar] [CrossRef] [Green Version]
  188. Broustas, C.G.; Hopkins, K.M.; Panigrahi, S.K.; Wang, L.; Virk, R.K.; Lieberman, H.B. RAD9A promotes metastatic phenotypes through transcriptional regulation of anterior gradient 2 (AGR2). Carcinogenesis 2018, 40, 164–172. [Google Scholar] [CrossRef]
  189. Cheng, C.K.; Chow, L.W.; Loo, W.T.; Chan, T.K.; Chan, V. The cell cycle checkpoint gene rad9 is a novel oncogene activated by 11q13 amplification and dna methylation in breast cancer. Cancer Res. 2005, 65, 8646–8654. [Google Scholar] [CrossRef] [Green Version]
  190. Galetzka, D.; Boeck, J.; Dittrich, M.; Sinizyn, O.; Ludwig, M.; Rossmann, H.; Spix, C.; Radsak, M.; Scholz-Kreisel, P.; Mirsch, J.; et al. Schmidberger HHypermethylation of RAD9A intron 2 in childhood cancer patients, leukemia and tumor cell lines suggest a role for oncogenic transformation. EXCLI J 2022, 21, 117–143. [Google Scholar]
  191. Lee, H.S.; Cho, S.-B.; Lee, H.E.; Kim, M.A.; Kim, J.H.; Park, D.J.; Kim, J.H.; Yang, H.-K.; Lee, B.L.; Kim, W.H. Protein expression profiling and molecular classification of gastric cancer by the tissue array method. Clin. Cancer Res. 2007, 13, 4154–4163. [Google Scholar] [CrossRef] [Green Version]
  192. Hopkins, K.M.; Auerbach, W.; Wang, X.Y.; Hande, M.P.; Hang, H.; Wolgemuth, D.J.; Joyner, A.L.; Lieberman, H.B. Deletion of mouse rad9 causes abnormal cellular responses to DNA damage, genomic instability, and embryonic lethality. Mol. Cell Biol. 2004, 24, 7235–7248. [Google Scholar] [CrossRef] [Green Version]
  193. Hu, Z.; Liu, Y.; Zhang, C.; Zhao, Y.; He, W.; Han, L.; Yang, L.; Hopkins, K.M.; Yang, X.; Lieberman, H.B.; et al. Targeted deletion of Rad9 in mouse skin keratinocytes enhances genotoxin-induced tumor development. Cancer Res. 2008, 68, 5552–5561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Mazur, D.J.; Perrino, F.W. Structure and expression of the TREX1 and TREX2 3′–> 5′ exonuclease genes. J. Biol. Chem. 2001, 276, 14718–14727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Höss, M.; Robins, P.; Naven, T.J.; Pappin, D.; Sgouros, J.; Lindahl, T. A human DNA editing enzyme homologous to the Escherichia coli DnaQ/MutD protein. EMBO J. 1999, 18, 3868–3875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Zhou, W.; Mohr, L.; Maciejowski, J.; Kranzusch, P.J. cGAS phase separation inhibits TREX1-mediated DNA degradation and enhances cytosolic DNA sensing. Mol. Cell 2021, 81, 739–755.e7. [Google Scholar] [CrossRef]
  197. Mohr, L.; Toufektchan, E.; von Morgen, P.; Chu, K.; Kapoor, A.; Maciejowski, J. ER-directed TREX1 limits cGAS activation at micronuclei. Mol. Cell 2021, 81, 724–738.e9. [Google Scholar] [CrossRef] [PubMed]
  198. Kumar, S.; Morrison, J.H.; Dingli, D.; Poeschla, E. HIV-1 Activation of innate immunity depends strongly on the intracellular level of TREX1 and sensing of incomplete reverse transcription products. J. Virol. 2018, 92, e00001-18. [Google Scholar] [CrossRef] [Green Version]
  199. Vanpouille-Box, C.; Alard, A.; Aryankalayil, M.J.; Sarfraz, Y.; Diamond, J.M.; Schneider, R.J.; Inghirami, G.; Coleman, C.N.; Formenti, S.C.; DeMaria, S. DNA exonuclease Trex1 regulates radiotherapy-induced tumour immunogenicity. Nat. Commun. 2017, 8, 15618. [Google Scholar] [CrossRef]
  200. Hasan, M.; Gonugunta, V.K.; Dobbs, N.; Ali, A.; Palchik, G.; Calvaruso, M.A.; DeBerardinis, R.J.; Yan, N. Chronic innate immune activation of TBK1 suppresses mTORC1 activity and dysregulates cellular metabolism. Proc. Natl. Acad. Sci. USA 2017, 114, 746–751. [Google Scholar] [CrossRef] [Green Version]
  201. Hasan, M.; Fermaintt, C.S.; Gao, N.; Sakai, T.; Miyazaki, T.; Jiang, S.; Li, Q.-Z.; Atkinson, J.P.; Morse, H.C.; Lehrman, M.A.; et al. Cytosolic nuclease TREX1 regulates Oligosaccharyl transferase activity independent of nuclease activity to suppress immune activation. Immunity 2015, 43, 463–474. [Google Scholar] [CrossRef] [Green Version]
  202. Lee-Kirsch, M.A.; Gong, M.; Chowdhury, D.; Senenko, L.; Engel, K.; Lee, Y.-A.; De Silva, U.; Bailey, S.L.; Witte, T.; Vyse, T.J.; et al. Mutations in the gene encoding the 3′-5′ DNA exonuclease TREX1 are associated with systemic lupus erythematosus. Nat. Genet. 2007, 39, 1065–1067. [Google Scholar] [CrossRef] [PubMed]
  203. Crow, Y.J.; Hayward, B.E.; Parmar, R.; Robins, P.; Leitch, A.; Ali, M.; Black, D.N.; van Bokhoven, H.; Brunner, H.G.; Hamel, B.C.; et al. Mutations in the gene encoding the 3′-5′ DNA exonuclease TREX1 cause Aicardi-Goutières syndrome at the AGS1 locus. Nat. Genet. 2006, 38, 917–920. [Google Scholar] [CrossRef] [PubMed]
  204. Simpson, S.R.; Hemphill, W.O.; Hudson, T.; Perrino, F.W. TREX1—Apex predator of cytosolic DNA metabolism. DNA Repair 2020, 94, 102894. [Google Scholar] [CrossRef]
  205. Yan, N. Immune diseases associated with TREX1 and STING dysfunction. J. Interf. Cytokine Res. 2017, 37, 198–206. [Google Scholar] [CrossRef] [PubMed]
  206. Morita, M.; Stamp, G.; Robins, P.; Dulic, A.; Rosewell, I.; Hrivnak, G.; Daly, G.; Lindahl, T.; Barnes, D.E. Gene-targeted mice lacking the Trex1 (DNase III) 3′-->5′ DNA exonuclease develop inflammatory myocarditis. Mol. Cell Biol. 2004, 24, 6719–6727. [Google Scholar] [CrossRef] [Green Version]
  207. Sakai, T.; Miyazaki, T.; Shin, D.-M.; Kim, Y.-S.; Qi, C.-F.; Fariss, R.; Munasinghe, J.; Wang, H.; Kovalchuk, A.L.; Kothari, P.H.; et al. DNase-active TREX1 frame-shift mutants induce serologic autoimmunity in mice. J. Autoimmun. 2017, 81, 13–23. [Google Scholar] [CrossRef] [Green Version]
  208. Grieves, J.L.; Fye, J.M.; Harvey, S.; Grayson, J.M.; Hollis, T.; Perrino, F.W. Exonuclease TREX1 degrades double-stranded DNA to prevent spontaneous lupus-like inflammatory disease. Proc. Natl. Acad. Sci. USA 2015, 112, 5117–5122. [Google Scholar] [CrossRef] [Green Version]
  209. Miyazaki, T.; Kim, Y.S.; Yoon, J.; Wang, H.; Suzuki, T.; Morse, H.C. The 3′-5′ DNA exonuclease TREX1 directly interacts with poly(ADP-ribose) polymerase-1 (PARP1) during the DNA damage response. J. Biol. Chem. 2014, 289, 32548–32558. [Google Scholar] [CrossRef] [Green Version]
  210. Yang, Y.-G.; Lindahl, T.; Barnes, D.E. Trex1 exonuclease degrades ssDNA to prevent chronic checkpoint activation and autoimmune disease. Cell 2007, 131, 873–886. [Google Scholar] [CrossRef] [Green Version]
  211. Maciejowski, J.; Chatzipli, A.; Dananberg, A.; Chu, K.; Toufektchan, E.; Klimczak, L.J.; Gordenin, D.A.; Campbell, P.J.; De Lange, T. APOBEC3-dependent kataegis and TREX1-driven chromothripsis during telomere crisis. Nat. Genet. 2020, 52, 884–890. [Google Scholar] [CrossRef]
  212. Han, G.; Tian, Y.; Duan, B.; Sheng, H.; Gao, H.; Huang, J. Association of nuclear annexin A1 with prognosis of patients with esophageal squamous cell carcinoma. Int. J. Clin. Exp. Pathol. 2014, 7, 751–759. [Google Scholar] [PubMed]
  213. Prati, B.; Abjaude, W.D.S.; Termini, L.; Morale, M.; Herbster, S.; Longatto-Filho, A.; Nunes, R.A.L.; Camacho, L.C.C.; Rabelo-Santos, S.H.; Zeferino, L.C.; et al. Three prime repair exonuclease 1 (TREX1) expression correlates with cervical cancer cells growth in vitro and disease progression in vivo. Sci. Rep. 2019, 9, 351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Tomicic, M.T.; Aasland, D.; Nikolova, T.; Kaina, B.; Christmann, M. Human three prime exonuclease TREX1 is induced by genotoxic stress and involved in protection of glioma and melanoma cells to anticancer drugs. Biochim. Biophys. Acta 2013, 1833, 1832–1843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Feng, J.; Lan, R.; Cai, G.; Lin, J. TREX1 suppression imparts cancer-stem-cell-like characteristics to CD133- osteosarcoma cells through the activation of E2F4 signaling. Int. J. Clin. Exp. Pathol. 2019, 12, 1134–1153. [Google Scholar]
  216. Erdal, E.; Haider, S.; Rehwinkel, J.; Harris, A.L.; McHugh, P.J. A prosurvival DNA damage-induced cytoplasmic interferon response is mediated by end resection factors and is limited by Trex1. Genes Dev. 2017, 31, 353–369. [Google Scholar] [CrossRef] [Green Version]
  217. Nader, G.P.D.F.; Agüera-Gonzalez, S.; Routet, F.; Gratia, M.; Maurin, M.; Cancila, V.; Cadart, C.; Palamidessi, A.; Ramos, R.N.; Roman, M.S.; et al. Compromised nuclear envelope integrity drives TREX1-dependent DNA damage and tumor cell invasion. Cell 2021, 184, 5230–5246.e22. [Google Scholar] [CrossRef]
  218. Garcia, F.A.d.O.; de Andrade, E.S.; Galvão, H.D.C.R.; Sábato, C.D.S.; Campacci, N.; de Paula, A.E.; Evangelista, A.F.; Santana, I.V.V.; Melendez, M.E.; Reis, R.M.; et al. New insights on familial colorectal cancer type X syndrome. Sci. Rep. 2022, 12, 2846. [Google Scholar] [CrossRef]
  219. Vanpouille-Box, C.; Formenti, S.C.; DeMaria, S. TREX1 dictates the immune fate of irradiated cancer cells. OncoImmunology 2017, 6, e1339857. [Google Scholar] [CrossRef] [Green Version]
  220. Ma, Z.; Xiong, Q.; Xia, H.; Liu, W.; Dai, S.; Cai, S.; Yan, X. Carboplatin activates the cGAS-STING pathway by upregulating the TREX-1 (three prime repair exonuclease 1) expression in human melanoma. Bioengineered 2021, 12, 6448–6458. [Google Scholar] [CrossRef]
  221. Wang, C.-J.; Lam, W.; Bussom, S.; Chang, H.-M.; Cheng, Y.-C. TREX1 acts in degrading damaged DNA from drug-treated tumor cells. DNA Repair 2009, 8, 1179–1189. [Google Scholar] [CrossRef] [Green Version]
  222. Hemphill, W.O.; Simpson, S.R.; Liu, M.; Salsbury, F.R.J.; Hollis, T.; Grayson, J.M.; Perrino, F.W. TREX1 as a novel immunotherapeutic target. Front. Immunol. 2021, 12, 660184. [Google Scholar] [CrossRef] [PubMed]
  223. Mazur, D.J.; Perrino, F.W. Identification and expression of the TREX1 and TREX2 cDNA sequences encoding mammalian 3′-–>5′ exonucleases. J. Biol. Chem. 1999, 274, 19655–19660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Parra, D.; Manils, J.; Castellana, B.; Viña-Vilaseca, A.; Morán-Salvador, E.; Vázquez-Villoldo, N.; Tarancón, G.; Borràs, M.; Sancho, S.; Benito, C.; et al. Increased susceptibility to skin carcinogenesis in TREX2 knockout mice. Cancer Res. 2009, 69, 6676–6684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Perrino, F.W.; Harvey, S.; McMillin, S.; Hollis, T. The human TREX2 3′–>5′-exonuclease structure suggests a mechanism for efficient nonprocessive DNA catalysis. J. Biol. Chem. 2005, 280, 15212–15218. [Google Scholar] [CrossRef] [Green Version]
  226. Manils, J.; Casas, E.; Viña-Vilaseca, A.; López-Cano, M.; Díez-Villanueva, A.; Gómez, D.; Marruecos, L.; Ferran, M.; Benito, C.; Perrino, F.W.; et al. The exonuclease Trex2 shapes psoriatic phenotype. J. Investig. Dermatol. 2016, 136, 2345–2355. [Google Scholar] [CrossRef] [Green Version]
  227. Manils, J.; Fischer, H.; Climent, J.; Casas, E.; García-Martínez, C.; Bas, J.; Sukseree, S.; Vavouri, T.; Ciruela, F.; de Anta, J.M.; et al. Double deficiency of Trex2 and DNase1L2 nucleases leads to accumulation of DNA in lingual cornifying keratinocytes without activating inflammatory responses. Sci. Rep. 2017, 7, 11902. [Google Scholar] [CrossRef] [Green Version]
  228. Weiss, T.; Wang, C.; Kang, X.; Zhao, H.; Gamo, M.E.; Starker, C.G.; Crisp, P.A.; Zhou, P.; Springer, N.M.; Voytas, D.F.; et al. Optimization of multiplexed CRISPR/Cas9 system for highly efficient genome editing in Setaria viridis. Plant J. 2020, 104, 828–838. [Google Scholar] [CrossRef]
  229. Chari, R.; Mali, P.; Moosburner, M.; Church, G.M. Unraveling CRISPR-Cas9 genome engineering parameters via a library-on-library approach. Nat. Methods 2015, 12, 823–826. [Google Scholar] [CrossRef] [Green Version]
  230. Bothmer, A.; Phadke, T.; Barrera, L.A.; Margulies, C.M.; Lee, C.S.; Buquicchio, F.; Moss, S.; Abdulkerim, H.S.; Selleck, W.; Jayaram, H.; et al. Characterization of the interplay between DNA repair and CRISPR/Cas9-induced DNA lesions at an endogenous locus. Nat. Commun. 2017, 8, 13905. [Google Scholar] [CrossRef]
  231. Yin, J.; Lu, R.; Xin, C.; Wang, Y.; Ling, X.; Li, D.; Zhang, W.; Liu, M.; Xie, W.; Kong, L.; et al. Cas9 exo-endonuclease eliminates chromosomal translocations during genome editing. Nat. Commun. 2022, 13, 1204. [Google Scholar] [CrossRef]
  232. Dumitrache, L.C.; Hu, L.; Hasty, P. TREX2 exonuclease defective cells exhibit double-strand breaks and chromosomal fragments but not Robertsonian translocations. Mutat. Res. Mol. Mech. Mutagen. 2009, 662, 84–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Chen, M.-J.; Dumitrache, L.C.; Wangsa, D.; Ma, S.-M.; Padilla-Nash, H.; Ried, T.; Hasty, P. Cisplatin depletes TREX2 and causes Robertsonian translocations as seen in TREX2 knockout cells. Cancer Res. 2007, 67, 9077–9083. [Google Scholar] [CrossRef] [Green Version]
  234. Ko, J.H.; Son, M.Y.; Zhou, Q.; Molnarova, L.; Song, L.; Mlcouskova, J.; Jekabsons, A.; Montagna, C.; Krejci, L.; Hasty, P. TREX2 Exonuclease causes spontaneous mutations and stress-induced replication fork defects in cells expressing RAD51K133A. Cell Rep. 2020, 33, 108543. [Google Scholar] [CrossRef] [PubMed]
  235. Hu, L.; Kim, T.M.; Son, M.Y.; Kim, S.-A.; Holland, C.L.; Tateishi, S.; Kim, N.H.; Yew, P.R.; Montagna, C.; Dumitrache, L.C.; et al. Two replication fork maintenance pathways fuse inverted repeats to rearrange chromosomes. Nature 2013, 501, 569–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Weigel, C.; Chaisaingmongkol, J.; Assenov, Y.; Kuhmann, C.; Winkler, V.; Santi, I.; Bogatyrova, O.; Kaucher, S.; Bermejo, J.L.; Leung, S.Y.; et al. DNA methylation at an enhancer of the three prime repair exonuclease 2 gene (TREX2) is linked to gene expression and survival in laryngeal cancer. Clin. Epigenetics 2019, 11, 67. [Google Scholar] [CrossRef]
  237. Song, D.; Zhang, D.; Chen, S.; Wu, J.; Hao, Q.; Zhao, L.; Ren, H.; Du, N. Identification and validation of prognosis-associated DNA repair gene signatures in colorectal cancer. Sci. Rep. 2022, 12, 6946. [Google Scholar] [CrossRef] [PubMed]
  238. Mukherjee, S.; Sinha, D.; Bhattacharya, S.; Srinivasan, K.; Abdisalaam, S.; Asaithamby, A. Werner syndrome protein and DNA replication. Int. J. Mol. Sci. 2018, 19, 3442. [Google Scholar] [CrossRef] [Green Version]
  239. Savva, C.; Sadiq, M.; Sheikh, O.; Karim, S.; Trivedi, S.; Green, A.R.; Rakha, E.A.; Madhusudan, S.; Arora, A. Werner syndrome protein expression in breast cancer. Clin. Breast Cancer 2020, 21, 57–73.e7. [Google Scholar] [CrossRef]
  240. Agrelo, R.; Cheng, W.-H.; Setien, F.; Ropero, S.; Espada, J.; Fraga, M.F.; Herranz, M.; Paz, M.F.; Sanchez-Cespedes, M.; Artiga, M.J.; et al. Epigenetic inactivation of the premature aging Werner syndrome gene in human cancer. Proc. Natl. Acad. Sci. USA 2006, 103, 8822–8827. [Google Scholar] [CrossRef] [Green Version]
  241. Chughtai, S.A.; Crundwell, M.C.; Cruickshank, N.R.; Affie, E.; Armstrong, S.; Knowles, M.A.; Takle, L.A.; Kuo, M.; Khan, N.; Phillips, S.M.; et al. Two novel regions of interstitial deletion on chromosome 8p in colorectal cancer. Oncogene 1999, 18, 657–665. [Google Scholar] [CrossRef] [Green Version]
  242. Armes, J.E.; Hammet, F.; de Silva, M.; Ciciulla, J.; Ramus, S.J.; Soo, W.-K.; Mahoney, A.; Yarovaya, N.; Henderson, M.A.; Gish, K.; et al. Candidate tumor-suppressor genes on chromosome arm 8p in early-onset and high-grade breast cancers. Oncogene 2004, 23, 5697–5702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Prorok, P.; Grin, I.; Matkarimov, B.; Ishchenko, A.; Laval, J.; Zharkov, D.; Saparbaev, M. Evolutionary origins of dna repair pathways: Role of oxygen catastrophe in the emergence of DNA glycosylases. Cells 2021, 10, 1591. [Google Scholar] [CrossRef] [PubMed]
  244. Friedberg, E.C.; McDaniel, L.D.; Schultz, R.A. The role of endogenous and exogenous DNA damage and mutagenesis. Curr. Opin. Genet. Dev. 2004, 14, 5–10. [Google Scholar] [CrossRef] [PubMed]
  245. Bermúdez-Guzmán, L. Pan-cancer analysis of non-oncogene addiction to DNA repair. Sci. Rep. 2021, 11, 23264. [Google Scholar] [CrossRef] [PubMed]
  246. Balian, A.; Hernandez, F.J. Nucleases as molecular targets for cancer diagnosis. Biomark. Res. 2021, 9, 86. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Exonucleases in DNA damage repair, apoptosis, and inflammation. Key exonuclease proteins in DNA damage repair, apoptosis, and inflammation processes are depicted. Exonuclease activity (5′–3′ and 3′–5′) is shown.
Figure 1. Exonucleases in DNA damage repair, apoptosis, and inflammation. Key exonuclease proteins in DNA damage repair, apoptosis, and inflammation processes are depicted. Exonuclease activity (5′–3′ and 3′–5′) is shown.
Cells 11 02157 g001
Figure 2. Exonuclease protein–protein interaction map. The interaction network of selected exonucleases was generated with STRING database v11.5 using basic settings selecting a physical subnetwork (the edges indicate that the proteins are part of a physical complex, although they may not directly interact) and medium confidence of 0.4. Proteins are represented as nodes, and lines indicate associations based on known functional interactions in humans. The network is significantly enriched in interactions (PPI enrichment p-value: 1.07e–13, FDR  <  0.05). All the proteins are included in the GO-term DNA metabolic process (in blue) (GO:0006259).
Figure 2. Exonuclease protein–protein interaction map. The interaction network of selected exonucleases was generated with STRING database v11.5 using basic settings selecting a physical subnetwork (the edges indicate that the proteins are part of a physical complex, although they may not directly interact) and medium confidence of 0.4. Proteins are represented as nodes, and lines indicate associations based on known functional interactions in humans. The network is significantly enriched in interactions (PPI enrichment p-value: 1.07e–13, FDR  <  0.05). All the proteins are included in the GO-term DNA metabolic process (in blue) (GO:0006259).
Cells 11 02157 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manils, J.; Marruecos, L.; Soler, C. Exonucleases: Degrading DNA to Deal with Genome Damage, Cell Death, Inflammation and Cancer. Cells 2022, 11, 2157. https://doi.org/10.3390/cells11142157

AMA Style

Manils J, Marruecos L, Soler C. Exonucleases: Degrading DNA to Deal with Genome Damage, Cell Death, Inflammation and Cancer. Cells. 2022; 11(14):2157. https://doi.org/10.3390/cells11142157

Chicago/Turabian Style

Manils, Joan, Laura Marruecos, and Concepció Soler. 2022. "Exonucleases: Degrading DNA to Deal with Genome Damage, Cell Death, Inflammation and Cancer" Cells 11, no. 14: 2157. https://doi.org/10.3390/cells11142157

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