Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Mar 1.
Published in final edited form as: DNA Repair (Amst). 2021 Jan 21;99:103049. doi: 10.1016/j.dnarep.2021.103049

The Multifaceted Roles of DNA Repair and Replication Proteins in Aging and Obesity

Alexandra M D’Amico 1, Karen M Vasquez 1,*
PMCID: PMC7941874  NIHMSID: NIHMS1671226  PMID: 33529944

Abstract

Efficient mechanisms for genomic maintenance (i.e., DNA repair and DNA replication) are crucial for cell survival. Aging and obesity can lead to the dysregulation of genomic maintenance proteins/pathways and are significant risk factors for the development of cancer, metabolic disorders, and other genetic diseases. Mutations in genes that code for proteins involved in DNA repair and DNA replication can also exacerbate aging- and obesity-related disorders and lead to the development of progeroid diseases. In this review, we will discuss the roles of various DNA repair and replication proteins in aging and obesity as well as investigate the possible mechanisms by which aging and obesity can lead to the dysregulation of these proteins and pathways.

Keywords: Aging, obesity, DNA repair, DNA replication, genetic instability

1. Introduction

Since the discovery of the double-helical structure of DNA over 50 years ago, an incredible effort has gone into understanding how genome stability is maintained. Thousands of DNA lesions occur per cell per day, resulting from both endogenous and exogenous sources. If left unrepaired or repaired in an error-generating fashion, then these lesions can lead to an accumulation of DNA damage and genomic instability. Multiple DNA repair pathways have evolved to remove various types of DNA damage from the genome. For example, exposure to ultraviolet light from the sun can lead to the formation of pyrimidine dimers, which are recognized and repaired by the nucleotide excision repair (NER) pathway[1]. Reactive oxygen species (ROS) generated from exogenous sources and metabolic processes within the cell can result in oxidative DNA damage that can be repaired by the base excision repair pathway (BER)[2]. DNA double-strand breaks (DSBs) caused by replication fork collapse or γ-irradiation, for example, are repaired by homologous recombination (HR) and non-homologous end-joining (NHEJ) pathways[3, 4]. Thus, the dysregulation or impairment of these DNA repair pathways can lead to an accumulation of DNA damage and/or mutations, resulting in genomic instability and the development of various genetic diseases.

Aging is a dynamic process by which various molecular processes decline over time, ultimately resulting in senescence and/or cell death. Common features associated with aging include graying hair, loss of skin plasticity, impaired organ function, and increased susceptibility to disease. Molecular hallmarks of aging include telomere shortening, early-onset of cellular senescence, increased DNA damage, decreased DNA repair, mitochondrial dysfunction, and loss of proteostasis (Figure 1)[5]. Mutations in several different DNA repair proteins can result in accelerated aging disorders (i.e., progeroid disorders), with the end result often being premature death. Many of these proteins and their involvement in preventing premature aging will be discussed in this review.

Figure 1.

Figure 1.

Common hallmarks in aging and obesity. Characteristics frequently observed in aging (left) and obesity (right) are illustrated above. Hallmarks common to both aging and obesity are displayed in the overlapping section of the diagram and are referred to in the text. ROS, reactive oxygen species.

A number of theories have been proposed to explain the aging process[6]. The “telomere theory of aging” suggests that telomeres, a DNA sequence repeat that caps chromosome ends to protect the genome, shorten with each round of DNA replication and upon reaching a certain length the cells are unable to continue replicating, resulting in senescence and cell death[7]. Another theory, the “free radical theory of aging”, suggests that free radicals produced by endogenous processes attack various macromolecules, including proteins and DNA, resulting in an accumulation of DNA damage that with time overwhelms the repair capacity of the cell, resulting in cell death (Figure 2)[8]. It is likely that both of these events, among others, contribute to the overall process of aging, and thus understanding the underlying mechanisms involved is paramount in preventing premature aging and age-associated diseases.

Figure 2.

Figure 2.

Schematic overview demonstrating how aging and obesity can lead to genetic instability. ROS, reactive oxygen species.

Another state that can influence genomic maintenance and stability is obesity. Obesity is most commonly characterized by an excessive increase in adipose tissue and chronic low-grade inflammation, which can lead to an accumulation of toxic by-products[9]. For example, individuals with obesity often have increased levels of ROS, which can oxidize DNA and lead to DNA damage[10]. Further, individuals with obesity have also been reported to have decreased levels of DNA repair, resulting in an accumulation of DNA damage, mutations, and genomic instability (Figure 2)[11].

In this review, we investigate how aging and obesity play roles in the dysregulation of DNA repair processes, as well as how defects in repair and replication proteins can exacerbate aging- and obesity-related diseases.

2. DNA Damage Repair

Genetic instability is a common underlying event in conditions of aging and obesity and plays a role in many associated disorders. Therefore, it is imperative that our cells have efficient mechanisms of DNA repair to ensure maintenance of genomic integrity. Several DNA repair processes have evolved to recognize and repair different types of DNA damage, including damaged or mismatched bases, single-stranded breaks (SSBs), and DSBs. In this section, we will discuss the roles of DNA repair proteins in aging- and obesity-related diseases and how defects in these pathways and proteins contribute to these diseases (see Table 1).

Table 1.

Roles and functions of DNA repair and replication proteins in aging- and obesity-related phenotypes.

Protein DNA Repair Pathway Role Phenotypes References
Pre-mature
Aging
Metabolic
Dysfunction
Cellular
Senescence
Telomere
Dysfunction
OGGI BER Removes 8-oxoG lesions [32,33,45,46,47,52,53,54,55]
NEIL Proteins BER Removes oxidized pyrimidines, oxidized cytosine derivatives, FapyA, 8-oxoA, and thymine glycol [56,58,59,67,68,69]
APE1 BER Recognizes AP site after removal of lesion and creates a nick for gap synthesis to occur [72,73]
POLB BER Removes 5'dRP during short-patch BER and fills in the nucleotide gap left behind by APE1 [74,75,76]
XRCC1 BER Complexes with LigaseIII to seal in the nick via ligase activity and complete the BER process [79]
ERCC1-XPF NER Performs DNA incision at the 5'-end of the lesion [84,85,86,90,91,97]
XPA NER Verifies lesion and recruits other repair factors [97,100]
XPB NER Subunit of TFIIH and involved in unwinding of double helix after damage recognition [97,103]
xpc NER Recognizes bulky lesion during global genome NER [97,101]
XPD NER Subunit of TFIIH and involved in unwinding of double helix after damage recognition [97,103]
XPG NER Performs DNA incision at the 3'-end of the lesion [97,102]
XPV NER Functions as a translesion synthesis polymerase during lesion bypass [97,105,106]
CSA NER Recruits other NER factors during TC-NER [107,108,110]
CSB NER Recruits other NER factors during TC-NER [107,108,111,113,114,115]
Ku70/80 NHEJ Recognizes and stabilizes broken DNA ends [153,154]
DNA-PKcs NHEJ Phosphorylates NHEJ factors and induces a conformational change to allow access of broken ends to NHEJ proteins [154]
Rad51 HR Binds ssDNA and facilitates strand invasion into the sister chromatid [171,172,173,176]
BRCA2 HR Mediates recruitment and assembly of Rad51 to ssDNA prior to strand invasion [172,173]
Rad54 HR Interacts with and guides Rad51-ssDNA complex along sister chromatid dsDNA during homology search [175]
WRN DNA Replication/DNA Repair Functions as a DNA helicase to unwind the DNA for various metabolic process [180,181,183,186,189,190,191,192,194]
BLM DNA Replication/DNA Repair Functions as a DNA helicase to unwind the DNA for various metabolic process [201,202,203,204,205]
FEN1 DNA Replication/BER Cleaves 5'-flap structures that arise during DNA replication and long-patch BER [217,218,222,223]
Twinkle Mitochondrial DNA Replication Functions as a mitochondrial DNA helicase to unwind the DNA at the replication fork [226,227,228,229]

2.1. Base Excision Repair (BER)

The BER pathway is a conserved and ubiquitous mechanism responsible predominately for the repair of single base, non-helix-distorting DNA damage (e.g., oxidative damage, deamination, alkylation, etc.). Initiated by various DNA glycosylases, the BER pathway is responsible for recognizing and repairing thousands of lesions per cell per day[2, 12]. Briefly, a lesion-specific glycosylase recognizes and removes the damaged base by cleaving the glycosidic bond between the damaged base and phosphate backbone, leaving an apurinic/apyrimidinic site (AP site). AP Endonuclease 1 (APE1) then makes an incision in the backbone generating 3’hydroxyl (3’OH) and 5’deoxyribosephosphate (5’dRP) ends[13]. In most cases, the 5’dRP moiety can be removed by polymerase β (pol β), which then uses the opposite DNA strand as a template to fill in the nucleotide gap. Finally, DNA ligase IIIα (LigIII) and X-ray cross-complementing protein 1 (XRCC1) form a complex to seal the remaining nick, thereby completing the short-patch BER pathway[14]. However, when pol β is unable to remove the 5’dRP (e.g., the moiety has undergone oxidation, reduction, etc.), then another sub-pathway of BER, long-patch BER, can process the damage. In long-patch BER, a single nucleotide is still inserted by pol β, but then polymerase delta/epsilon (pol δ/ε) add additional bases following the newly inserted base. The addition of several bases by pol δ/ε displaces downstream nucleotides, creating a 5’flap containing the 5’dRP, which is recognized and processed by flap endonuclease 1 (FEN1) and proliferating cell nuclear antigen (PCNA), leaving a nick in the backbone that is sealed by DNA ligase I (LigI) and PCNA[15-17]. An overview of this pathway is summarized in Figure 3. With over 20,000 base lesions occurring in each cell per day, BER is crucial for maintaining genomic integrity, and defects in this pathway can result in the development of several diseases, including cancer[18-21]. In fact, mouse models deficient in the BER pathway, particularly via pol β and APE1 mutations, are either embryonically lethal or die shortly after birth[22-24]. In this section, we will discuss several BER-related enzymes and their roles in aging and obesity.

Figure 3.

Figure 3.

Schematic overview of DNA repair proteins and pathways discussed in this review.

2.1.1. 8-Oxoguanine DNA Glycosylase (OGG1)

OGG1 is a lesion-specific glycosylase in the BER pathway primarily responsible for the removal of 7,8-dihydro-8-oxoguanine (8-oxoG), a common lesion formed by ROS. If left unrepaired, this mutated base can mispair with adenine during DNA replication, generating a G:C to T:A transversion[25-28]. 8-oxoG is one of the most common oxidative lesions formed and, given its mispairing tendencies, is highly mutagenic[29, 30]. OGG1 is localized to both the nucleus and mitochondria. It has been reported that accumulation of mitochondrial DNA damage can lead to increased oxidative stress, mitochondrial dysfunction, and reduced insulin sensitivity[31]. Several studies have reported that mice deficient in OGG1 demonstrated a lower tolerance for cellular oxidative stress and were more prone to developing metabolic disorders[32, 33]. In this regard, OGG1-deficient mice displayed significant weight gain when fed a high-fat diet (HFD) and showed signs of fatty liver, including increased fat mass, lipid accumulation, triglyceride content, and insulin resistance, all signs of metabolic dysfunction[32]. These observations may be explained, at least in part, by reduced hepatic glycogen content, increased glycolytic gene expression, and increased respiratory exchange ratios in OGG1-deficient mice compared with wild-type (WT) mice when fed HFDs, all of which indicated decreased hepatic fatty acid oxidation[32]. These results suggested that reliance on fatty acids as fuel sources had decreased, while reliance on glycogen as a fuel source increased in the OGG1-deficient mice.

Interestingly, Scheffler et al. (2018) reported that OGG1-deficient mice on chow diets demonstrated increased expression of genes involved in gluconeogenesis in hepatic mitochondria compared to WT mice, contrary to the results from the mice on high-fat diets, where the expression of glycolytic genes was increased compared to WT mice[32, 33]. This discrepancy in gene expression could indicate different roles for OGG1 in regulating mitochondrial function during different energy states. In contrast to both of these studies, which demonstrated altered mitochondrial respiration and dysfunction, Stuart et al. (2005) found that mitochondrial respiration remained unaltered in OGG1-deficient mice[34]. It should be noted, however, that this group used mouse tissues pooled from various age groups, while the other studies used age-matched tissues, which may have impacted the results[32-34]. It has been well established that mitochondrial dysfunction is a characteristic of aging[35]. This could explain the lack of observable differences in mitochondrial respiration in OGG1-deficient and WT mice, further emphasizing the importance of sources of variation when utilizing mouse models.

These observed changes in OGG1-deficient mice are not limited to the liver, as another study reported that skeletal muscle tissue also presented with lipid accumulation, altered tissue metabolism, and decreased muscle function[36]. Several groups have also found polymorphisms in OGG1 in human patients that correlated with reduced insulin sensitivity, type 2 diabetes mellitus (T2DM), and obesity[37-39]. Additionally, the Lloyd and Sampath groups have reported in a collaborative study that overexpression of OGG1 actually protected mice against diet-induced obesity and suggested a role for OGG1 in maintaining mitochondrial energetics in white adipose tissue[40]. Together, these studies provide strong evidence for a role of OGG1 in modulating energy homeostasis on a whole-body level.

OGG1 has also been shown to be a regulator of age-related disorders. The “free-radical theory of aging” postulates that aging is largely the result of an accumulation of oxidative damage within the cell[41]. Consistent with this postulate, studies have shown that cellular senescence, a hallmark of aging, can be stimulated by high levels of oxidative stress[5, 42-44]. As one of the main enzymes involved in the repair of oxidative lesions, it can be hypothesized that OGG1 plays an important role in regulating the onset of aging and age-related diseases by facilitating the removal of the DNA damage caused by high levels of oxidative stress, thereby preventing the onset of cellular senescence. Indeed, several studies using various animal models have demonstrated increased levels of 8-oxoG and decreased levels of OGG1 with age[45-47]. Notably, many publications have reported that mitochondrial DNA, and not nuclear DNA, is where most of the oxidative damage is detected[28, 45, 47, 48]. Among many factors, this could be a result of: (1) nuclear DNA damage having a higher priority for DNA damage repair leading to lower levels of detected DNA damage compared to the mitochondria, and/or (2) the highly oxidative environment of the mitochondria, which could increase the susceptibility of mitochondrial DNA to oxidative damage compared to nuclear DNA. Regardless, further studies are required to address the mechanisms responsible for decreased 8-oxoG repair capacity in aging cells.

Another established theory of aging is the “telomeric theory of aging”[49]. Telomeres are comprised of DNA repeat sequences, (TTAGGG)n, that act as caps to chromosome ends to prevent DNA damage responses (DDR) from recognizing them as broken DNA. Telomere shortening is a hallmark of aging and can lead to cellular senescence and apoptosis[5, 7]. Telomeric sequences are rich in guanines, making these regions susceptible to oxidative damage. Thus, it has been thought that one of the primary sources of telomere shortening in cells is oxidative stress[50, 51]. Rhee et al. (2011) reported that telomeric DNA repeats demonstrated higher levels of oxidized guanines than non-telomeric minisatellite TG repeats[52, 53]. Due to the repetitive nature of telomeric DNA sequences, studies have shown that telomeric regions are able to form alternative DNA structures, e.g., G-quadruplex DNA structures, which are formed in guanine-rich regions and susceptible to oxidative damage[54]. Because of this propensity for oxidative damage, DNA repair mechanisms are crucial in maintaining telomere integrity. In support of this, Wang et al. (2010) demonstrated that aged OGG1-deficient mice and OGG1-deficient mouse embryonic fibroblasts (MEFs) cultured in highly oxidative environments (i.e., 20% oxygen) accumulated higher levels of oxidized guanine in telomeric DNA than young mice or MEFs under environments of lower oxidation (i.e., 3% oxygen)[50]. In addition, OGG1 depletion in Saccharomyces cerevisiae resulted in significantly higher levels of oxidized guanine residues in telomeric DNA sequences, as well as altered telomere lengths[55]. These data indicated that OGG1 plays a role in maintaining telomere homeostasis and may act to prevent the premature onset of cellular senescence associated with aging. Regardless of which theory of aging is being considered, it is evident that OGG1 plays a crucial role in maintaining genome integrity and preventing the onset of several hallmarks of aging, including mitochondrial dysfunction, telomere attrition, and genomic instability.

2.1.2. Nei Endonuclease VIII-like (NEIL) Proteins

Belonging to the Nei endonuclease VIII-like (NEIL) family of glycosylases, the NEIL proteins, similar to OGG1, are involved in the removal of damaged bases to initiate BER. However, unlike OGG1, the NEIL proteins are able to recognize a broader range of damaged substrates. Currently, there are three reported NEIL protein glycosylases: NEIL1, NEIL2, and NEIL3. NEIL1 is primarily involved in the removal of oxidized pyrimidines, such as thymine glycol and 5’hydroxyuracil, in concert with the replication fork machinery before the DNA is synthesized by polymerases. NEIL2, on the other hand, preferentially removes oxidized cytosine derivatives, particularly in mismatched double-stranded DNA such as during transcription-coupled DNA repair. Less is known about NEIL3, but studies have shown that it can remove a wide variety of lesions in single-stranded DNA, including 4,6-diamino-5-formamidopyrimidine (FapyA), 7,8-dihydro-8-oxoadenine (8-oxoA), and thymine glycol, and that it can also remove oxidized bases from G-quadruplex structures[56, 57]. Due to their roles in BER, and the role of BER in regulating oxidative stress, it is thought that, like OGG1, defects in the NEIL glycosylases may lead to metabolic dysfunction and obesity. Indeed, NEIL1-deficient mice have been reported to be prone to obesity, hepatic lipid accumulation, hepatic inflammation, hyperleptinemia, hyperinsulinemia, and hyperlipidemia, all of which are indicative of metabolic dysfunction[58, 59]. In regard to its role in DNA damage repair, NEIL1-deficient mice also have increased mitochondrial DNA damage and deletions[58].

When placed on HFD, NEIL1-deficient mice displayed similar but heightened metabolic stress and dysfunction compared to mice on chow diets, including the addition of impaired glucose clearance, decreased mitochondrial content, and decreased lean body mass[59]. It is interesting to note that while many similar outcomes were observed in OGG1-deficient mice, alterations in lean body mass had not been reported, suggesting that glycosylase-specific activities in addition to disruptions in the BER pathway likely contribute to metabolic dysregulation. Additionally, polymorphic variants of NEIL1 resulting in defective glycosylase activity have been reported in human cancer patients; however, susceptibility of these patients to metabolic disorders due to these variants has not been reported[60, 61].

Unlike NEIL1, very few studies have reported on the potential roles of NEIL2 or NEIL3 in energy balance and metabolic regulation. From a limited number of studies, it was observed that NEIL3 was involved in the regulation of lipid metabolism in atherosclerosis-prone mice, and that several NEIL3 polymorphisms were associated with type 2 diabetes[62, 63]. Although the roles of NEIL glycosylases in obesity and metabolic syndrome have not been extensively studied, these reports provide a solid foundation for further research into the underlying mechanisms imparted by the NEIL proteins on metabolic homeostasis.

Requiring nearly 20 percent of the body’s oxygen supply, the brain occupies an oxygen-rich environment and can be prone to high levels of oxidative stress and increased susceptibility to oxidative DNA damage. Therefore, it is imperative that DNA repair mechanisms be rapid and efficient in their recognition and removal of oxidative lesions in the brain. It has been observed that with age, the levels of DNA damage increase in the brain. But whether this is due to increased accumulation of damage or decreased repair capacity, or a combination thereof, remains unclear. DNA glycosylases play major roles in the repair of oxidative lesions via their recognition and removal of oxidized bases. Reduced activity of these proteins could lead to inefficient DNA repair and a subsequent increase in the levels of DNA lesions remaining in the genome, resulting in genetic instability. NEIL1 and NEIL2 mRNAs are widely distributed throughout the brain, supporting a role for BER in maintaining genomic integrity in the brain[64]. Interestingly, studies have shown that NEIL activity in various regions of the brain does not decline with age, as one might expect; in fact, NEIL activity was shown to peak in middle-aged mice before tapering down in older mice with levels comparable to young and adult-aged mice[65]. Consistent with this, Rolseth et al. (2008) demonstrated that NEIL1 mRNA in the brain was increased in 12-month-old mice compared to 1 month-old mice[64]. In a NEIL1-deficient mouse model, Canugovi et al. (2012) reported that NEIL1 was crucial for memory retention, as mice deficient in NEIL1 consistently had lower performance levels in Morris water maze behavioral tests compared to WT mice[66]. These mice also exhibited increased apoptosis, brain damage, decreased DNA repair capacity, and decreased motor function after ischemic stroke, suggesting that NEIL1 may act to protect the brain against damage induced by severe stress. These results suggested that NEIL proteins, and NEIL1 in particular, played crucial roles in maintaining genomic integrity in the brain. Thus, additional investigations into the mechanisms of their regulation and function in the brain with age are warranted.

The NEIL proteins have also been reported to be involved in the repair of telomeric DNA damage. As a hallmark of aging, telomere integrity is critical in preventing the onset of cellular senescence and apoptosis. Zhou et al. (2013) determined that NEIL1 and NEIL3 removed lesions in G-quadruplex-forming sequences with a preference for lesions in telomeric DNA sequences. Furthermore, NEIL3 was reported to be recruited to telomeres during S phase in human cells via interactions with telomeric repeat factor 1 (TRF1) to remove oxidative DNA lesions prior to cell division, such that loss of NEIL3 resulted in telomere dysfunction[56, 67]. Although NEIL2 did not demonstrate any activity on telomeric four-repeat quadruplex DNA (TTAGGG)4, NEIL2 did remove hydantoins from telomeric five-repeat quadruplex DNA (TTAGGG)5[68]. Moreover, Chakraborty et al. (2015) reported that embryonic fibroblasts from NEIL2-deficient mice had greater telomere instability than the WT mice, suggesting that NEIL2, like NEIL1 and NEIL3, plays some role in telomere stability and maintenance[69]. This is further supported by evidence that NEIL2 processes lesions within D-loops, which have been reported to form at telomeres[70].

2.1.3. Other Proteins Involved in Base Excision Repair

Glycosylases, such as those mentioned above, initiate BER via the removal of damaged bases leaving AP sites that are processed by a network of proteins, including APE1, Polβ, XRCC1, etc. APE1, also commonly referred to as reduction-oxidation factor 1 (REF-1), recognizes the AP site and incises the DNA backbone so that gap synthesis can occur. APE1 is a unique protein in the BER pathway as it was independently identified as a redox factor and a DNA repair protein, hence the two commonly used names REF-1 and APE1, respectively[71]. Polβ, a well-studied BER protein, acts to remove the 5’dRP moiety left behind by APE1 and fills in the nucleotide gap generated by the removal of the damaged base. Mutations in Polβ are common in many different types of cancer, including colon, breast, gastric, and lung cancer. Indeed, results from small-scale studies suggest that Polβ mutants are present in at least 30% of human tumors, a staggering percentage attributed to just one protein[18]. After Polβ performs gap synthesis, the XRCC1-LigIII complex then seals the nick via DNA ligase activity, thus completing the repair process.

Of the proteins involved in BER, only the DNA glycosylases have been thoroughly studied for their roles in obesity and metabolic syndrome. While the BER pathway itself is critical for repairing oxidative damage, particularly in high states of oxidative stress such as obesity, it has yet to be determined whether deficiency of proteins downstream of DNA glycosylases contribute to obesity or metabolic phenotypes. However, complete knockout of APE1 results in embryonic lethality, suggesting that one or more of its roles is essential for embryonic development[22]. Moreover, APE1 knockdown in human primary cells and tamoxifen-induced knockout of APE1 in mice resulted in induced telomeric DDR, telomere shortening, and cellular senescence, suggesting a role for APE1 in preventing aging-related phenotypes[72, 73].

Similar to APE1, studies have demonstrated that mice harboring mutations resulting in the inactivation of Polβ are not viable. For example, Gu et al. (1994) found that Polβ-null mice died during embryogenesis, sometime after embryonic day 10.5 (E10.5), indicating that Polβ knockout was embryonically lethal[23]. Sugo et al. (2000), however, found that Polβ-null mice did not die during embryogenesis, but rather shortly after birth, likely due to a respiratory defect[24]. This discrepancy could be due to the nature of the Polβ gene mutation resulting in the Polβ-null phenotype. Another possibility is that because the mice died so soon after birth, the mother disposed of the deceased pups before they were noted. In any case, it is apparent that Polβ is critical for early-life development. Interestingly, mice heterozygous for Polβ showed increased chromosomal abnormalities, increased single-stranded DNA breaks, and appeared to die at a faster rate than WT mice[74, 75]. Additionally, loss of Polβ induced cellular senescence in mouse primary cells[76]. These studies suggested that, like APE1, Polβ plays a crucial role in the onset of aging.

Although no direct studies have been conducted to investigate the mechanisms linking XRCC1 and obesity, correlation studies in humans have determined that polymorphisms in XRCC1 are associated with obesity and Type 2 diabetes[77, 78]. Further, XRCC1 deficiency in testis of mice induced oxidative stress, mitochondrial injury, and mitochondrial dysfunction[79]. Together, these results warrant further studies into the potential role(s) of XRCC1 in obesity and metabolic syndrome. Like APE1 and Polβ, deletion of both alleles of XRCC1 is embryonically lethal[80]. The observation that complete knockouts of a number of proteins in the BER pathway resulted in non-viable mice demonstrated the importance of BER in maintaining genomic integrity during early life development. Unlike APE1 and Polβ, however, no obvious links between XRCC1 and aging have been reported. In fact, mice heterozygous for XRCC1 have the same median life span as that of WT mice[81]. This is an interesting finding, since it is known that deficiencies in APE1 and Polβ resulted in several hallmarks of aging (i.e., genetic instability, senescence, telomere length alterations, etc.), and that these proteins are all crucial for functional BER. That XRCC1 deficiency does not result in aging phenotypes may suggest alternative mechanisms for completing the repair pathway in the absence of this scaffold protein.

2.2. Nucleotide Excision Repair (NER)

NER recognizes and removes bulky, helix-distorting lesions within the DNA, typically caused by UV radiation or chemical agents (e.g., Benzo(a)pyrene, cisplatin, etc.). There are two sub-pathways of NER, transcription-coupled NER (TC-NER) and global genome NER (GG-NER). The primary difference between the two sub-pathways lies in how they recognize the DNA damage. In TC-NER, stalling of RNA polymerase II during transcription leads to the recruitment of TC-NER specific enzymes, including Cockayne Syndrome groups A and B (CSA and CSB, respectively), which act to recruit other factors required for NER. In GG-NER, the damage is recognized by the distortions induced in the double helix by the damage via UV-damaged DNA-binding protein (UV-DDB) and Xeroderma pigmentosum complementation group C (XPC). From this point onward, the repair process is similar for both sub-pathways. Briefly, XPA is recruited to verify the damage as RPA binds the undamaged strand. Transcription Factor II H (TFIIH) is then recruited to assist in the formation of an open complex that is a substrate for the endonucleases, ERCC1-XPF and XPG, which make incisions at the 5’ and 3’ ends of the damage, respectively, resulting in an excised DNA fragment containing the lesion. Replication factor C (RFC), PCNA, Polε/δ, and DNA ligase are then recruited to fill in the gap left behind, completing the repair process (Figure 3)[82], There are a limited number of studies that have investigated obesity and NER; however, numerous studies have determined that defects in NER result in several progeroid syndromes (i.e., premature aging).

2.2.1. Excision Repair Cross-Complementation Group1-Xeroderma Pigmentosum Group F (ERCC1-XPF)

ERCC1-XPF is a heterodimeric endonuclease that is involved in a number of DNA repair processes, including its essential role in NER. Within the NER pathway, ERCC1-XPF acts by making an incision at the 5’ end of the bulky lesion, contributing in the excision and removal of the lesion from the DNA[83]. Though rare, mutations in the genes encoding XPF and ERCC1 have been reported to lead to cellular senescence and accelerated aging[84-86]. To date, only two patients have been reported to contain non-synonymous mutations in ERCC1, demonstrating the lethality of variants in this gene[87, 88]. Both patients had impaired NER capabilities, severe developmental symptoms, and decreased lifespan. ERCC1 knockout mice are viable for a period of time, but experience significance growth retardation, liver failure, and die around weaning age[89]. ERCC1 deficient mice, with one allele deleted and the other hypomorphic, are viable for a longer period of time and demonstrate phenotypes of accelerated aging, making them a valuable model for aging studies[90]. ERCC1-XPF deficiency resulted in accumulated DNA damage, enhanced senescence-associated secretory phenotype (SASP) factors, cellular senescence, and apoptosis in human primary cells and mouse tissues[85, 90, 91]. Almost counterintuitively, however, ERCC1-XPF has been reported to be a negative mediator in telomere length maintenance, facilitating telomere shortening via TRF2-associated mechanisms[92, 93]. This is interesting, as telomere shortening is a hallmark in aging and, as described previously, deficiency in ERCC1-XPF leads to senescence and accelerated aging. This may suggest that the mechanism(s) by which ERCC1-XPF deficiency leads to accelerated aging is independent of telomere shortening. Further studies investigating ERCC1-XPF and telomere maintenance are warranted to shed more light on this intriguing phenomenon.

Along with an accumulation of senescent cells, ERCC1-XPF-deficient mice also experience increased oxidative DNA damage and susceptibility to lipid peroxidation (LPO)-induced DNA damage[94, 95]. Dietary fats, particularly polyunsaturated fatty acids (PUFA), are a source for endogenous LPO, and LPO by-products have been reported to be increased in obese individuals[9]. When placed on diets high in PUFA, ERCC1-XPF-deficient mice showed decreased maximum lifespan and age-related degenerative markers in their livers and kidneys[95]. Interestingly, while mice on the PUFA diet had increased glycogen storage, there was no weight gain when compared to mice on the control diet, suggesting that alterations in metabolic homeostasis may be independent of caloric intake. Additionally, Karakasilioti et al. (2013) demonstrated that ERCC1-XPF deficiency induced lipodystrophy, which can lead to further metabolic complications and abnormalities, further supporting a role for ERCC1-XPF in metabolic regulation[96].

2.2.2. Xeroderma Pigmentosum (XP) Complementation Proteins

A total of eight XP complementation groups have been identified thus far, ranging from XPA to XPG and XPV. Mutations in these genes can result in the autosomal recessive disease XP, a skin condition that causes the patient to be highly sensitive to UV irradiation, increasing susceptibility to skin cancer, and premature aging. Conditions of XP range from mild to severe symptoms, depending on the mutation[97]. Mutations in the first seven XP groups, XPA-XPG, often lead to defects in the NER mechanism. The eighth gene, XPV, encodes for the translesion synthesis protein, polymerase eta (polη)[98]. The XP proteins have diverse roles in DNA repair ranging from damage recognition, validation, DNA excision, and lesion bypass during DNA replication[99]. While their roles are diverse in nature, they are all central in preventing DNA damage-induced genetic instability.

As mutations in these genes lead to UV sensitivity and premature aging, it stands that efficient repair of UV-induced DNA damage is critical in preventing the onset of premature aging. Indeed, Fang et al. (2014) found that across several species, including mice, rats, and humans, XPA deficiency resulted in mitochondrial dysfunction and degradation, a prominent hallmark of aging[100]. Additionally, XPC-deficient MEFs displayed increased telomere fragility compared to WT MEFs, and upon chronic UV exposure had increased telomere shortening compared to WT[101]. Harada et al. (1999) observed that mice deficient in XPG failed to mature properly after birth and underwent premature death around weaning age. They also observed that primary cells from these mice displayed an accumulation of p53 and premature senescence[102]. Deficiency in the NER DNA helicases, XPB or XPD, subunits of THIIH, increased the sensitivity of cells to oxidative DNA damage, particularly at telomeres, and resulted in telomere attrition[103]. This is an interesting outcome, as it supports the hypothesis that UV-induced DNA damage alone does not account for the various phenotypes observed in XP patients, including effects seen on organs that are not exposed to UV irradiation. Together, these studies support the idea that efficient repair of DNA damage is crucial to prevent the premature onset of aging. However, further studies are needed to uncover the underlying mechanisms of how these proteins are involved in preventing premature aging, particularly their roles in telomere maintenance, as deficiencies in several of these proteins result in telomere attrition.

Because XP patients have a predisposition for cancer development and premature aging, much of the published work investigating mechanisms of XP proteins focus on their roles in carcinogenesis and aging; very few studies have evaluated the potential role(s) for the XP proteins in metabolic homeostasis. XPB, a subunit of TFIIH that is canonically involved in the unwinding of the DNA double helix after damage recognition, was found to interact with the BER protein OGG1 in the hippocampus of cholesterol-fed rabbits[104]. As this interaction did not occur in rabbits on a normal diet, it supported the idea of cross-talk between DNA repair pathways in response to cholesterol-induced lipid peroxidation and oxidative stress, a phenomenon also observed during obesity[104].

Interestingly, in another studying investigating polη-deficiency, mice deficient in polη displayed increased weight gain with age compared to WT littermate mice independent of caloric intake or water consumption[105]. Upon further investigation, the mice also displayed increased body fat percentage, hyperinsulinemia, reduced glucose tolerance, hepatic steatosis, and adipocyte macrophage infiltration when compared to WT mice, all of which are characteristics of obesity and metabolic syndrome. Additionally, polη-deficient mice showed increased levels of DNA damage and cellular senescence; interestingly, the cellular senescence was detected as early as 4 weeks of age, prior to the onset of obesity in the mice[105]. Cellular senescence can lead to adipocyte hypertrophy and lipid accumulation, two primary characteristics of obesity[106]. This suggests that DNA damage-induced senescence may have played a causative role in the onset of obesity in polη-deficient mice. Altogether, this study suggests a role for polη in regulating metabolic homeostasis, likely via its function in lesion bypass to maintain genomic stability and prevent DNA damage-induced senescence, providing further evidence of a role for genome integrity in obesity and metabolic syndrome. However, additional studies are needed to corroborate these findings and further explore the underlying mechanisms linking DNA damage, repair, and obesity.

2.2.3. Cockayne Syndrome (CS) Proteins

Similar to XP, CS is another progeroid disorder that occurs in individuals lacking either CSA or CSB, and is characterized by short stature, microcephaly, neurodegeneration, UV sensitivity, and accelerated aging[107, 108]. Deficiency of CSA or CSB can lead to defects in TC-NER, which is crucial for repair of DNA damage in actively transcribed regions. After DNA damage recognition by transcription polymerase stalling, CSA and CSB are required for recruitment of NER factors, such as TFIIH, as well as proteins for transcription restart after the damage has been repaired[109]. Due to this, defects in CSA and CSB, and therefore defects in TC-NER, can lead to transcription blockage and persistent DNA damage, resulting in genomic instability and premature cell death. Human primary keratinocytes deficient in CSA displayed abnormal secretions of SASPs, an abundance of oxidative damage, and cellular senescence[110]. Re-expression of WT CSA rescued these functions, indicating a protective role for CSA in maintaining genomic integrity and cellular function to prevent the onset of premature aging.

CSB expression has been reported to decline in human cells and mouse brain tissue with age; additionally, CSB-deficient cells experience heterochromatin loss, a phenomenon typically seen in aging cells[111, 112]. Chrochemore et al. (2019) reported that downregulation of CSB was an early event in replication-dependent senescence and that depletion of CSB induced p21-dependent cellular senescence[111]. In addition to preventing heterochromatin loss and cellular senescence, CSB appears to play a role in telomere maintenance and stability. It has been reported that CSB may act in the suppression of ROS-induced DNA breaks at telomeres in an R-loop-dependent mechanism[113]. Indeed, Tan et al. (2020) demonstrated that cells deficient in CSB experienced greater telomere aberrations than WT cells, particularly upon ROS-induced oxidative damage. During replication stress, CSB has also been reported to recruit HR repair proteins to telomeres to prevent fork collapse and telomere fragility[114]. Additionally, Batenburg et al. (2012) demonstrated a role for CSB in maintaining telomere length and stability in a TRF2-dependent manner, where CSB was required for the suppression of telomere doublets and telomere dysfunction-induced foci (TIFs), both of which result in telomere instability[115]. Together, these results indicated that CSB plays a role in preventing the onset of aging through various mechanisms, although the roles of CSB in these pathways remain unclear.

2.3. Mismatch Repair (MMR)

The MMR pathway is primarily responsible for the recognition and repair of nucleotide misincorporations generated during DNA synthesis, resulting in either base-base mismatches or small insertion or deletion loops[116]. The functions of and steps involved in MMR have been reviewed extensively[116-118]. Briefly, the misincorporated base is recognized by either the MutS Homologue 2 (MSH2)/MutS Homologue 6 (MSH6) complex (MutSα) or the MSH2/MutS Homologue 3 (MSH3) complex (MutSβ), depending on the mismatch or loop structure. These complexes then recruit other factors, including MutL Homologue 1 (MLH1), PMS1 Homologue 2 (PMS2), and Exonuclease 1 (EXO1) for the removal of both the mismatched base and/or loop and several bases downstream of the damage. The repair factors PCNA, RFC, RPA, Polδ/ε, and DNA ligase then work together to fill in the gap to complete the process. A number of factors can influence the rate of nucleotide misincorporation, including the fidelity of the DNA polymerase (Figure 3). Replication polymerases typically have a high fidelity, generating few errors; however, translesion polymerases often have lower fidelity, generating mismatches more frequently than replication polymerases[116]. Similar to the other DNA repair pathways, MMR is critical for maintaining genomic stability and erroneous MMR can have deleterious consequences, including the generation of a mutator phenotype[119, 120]. Cells are constantly undergoing DNA replication and without MMR, mismatches generated by both high and low fidelity polymerases will accumulate, resulting in increased mutations, cellular senescence, and cell death. Defects in MMR have also been linked to several forms of cancer; in particular, MMR defects are frequently observed in patients with colorectal cancer (CRC)[121]. Obesity is recognized as a high-risk factor for the development of CRC[122]; however, the relationship between obesity and MMR in CRC is not well understood.

2.3.1. MutL Homologue 1 (MLH1)

MLH1 forms a heterodimer with PMS2 that functions primarily in the MMR pathway. This complex, known as MutLα, acts as a validation step, improving the recognition of the MutSα or MutSβ complexes to mismatched or looped DNA, respectively. MutLα also possesses cryptic endonuclease activity, which is thought to be required for the removal of the mismatched base via EXO1[116]. The lack of redundant binding partners for the formation of MutLα suggests that loss of MLH1 would hamper the MMR process, preventing the repair of mismatched or looped DNA[123]. Indeed, MLH1-deficient mice generated by Edelmann, et al. (1996) were replication-error positive (RER+) and demonstrated increased microsatellite instability and decreased MMR activity[124]. Additionally, both male and female MLH1-deficient mice were sterile due to meiotic disruption and apoptotic loss of germ cells. With an impaired MMR pathway, it was likely that the MLH1-deficient mouse germ cells accumulated too much DNA damage during meiosis, thereby resulting in apoptosis during the meiotic checkpoint. Further emphasizing the importance of MLH1 in the maintenance of genome integrity, Hegan et al. (2006) generated MLH1 knockout mice containing a recoverable mutation reporter and demonstrated that MLH1-deficient mice showed significantly higher mutation frequencies in skin and colon tissues than their WT counterparts[125].

Concurrent with genomic instability representing a hallmark of aging, several groups have reported associations between declining MMR activity and age[126-129]. Among the first to propose a relationship between MLH1 and age, Kim et al. (2006) analyzed and compared the coding region of MLH1 in individuals over 100-years old to a control group of ~50-years old and found that the centenarian population had an increased MLH1 haplotype frequency compared to the control group[127]. Although no potential mechanisms were proposed for this association, the outcome led to the idea that perhaps a stabilization of genome maintenance enzymes, particularly MMR enzymes, could enhance longevity. Likewise, a decrease in repair activity could lead to reduced longevity due to increased genomic instability. In support of this, Edelman et al. (1999) reported striking longevity differences between mice with MLH1 heterozygous genotypes and MLH1 null genotypes compared to WT littermates, with ~50% of the mice perishing at 18 months and 7 months, respectively[130]. Later, Kenyon et al. (2012) demonstrated decreased gene expression and protein levels of MLH1 that correlated with increased microsatellite instability events with age in hematopoietic stem cells[126]. The decrease in gene expression was attributed to epigenetic regulation, as the promoter region of MLH1 was increasingly methylated with age. This was an interesting finding, as it provided evidence for a proposed mechanism by which age-induced MMR deficiency could lead to increased genomic instability via transcriptional suppression of a specific DNA repair enzyme. While there is a foundation of evidence to support a role for MLH1 in aging, further studies are warranted, particularly in the context of animal models to investigate the underlying mechanisms of age-associated suppression of MLH1, and how this suppression could further exacerbate aging phenotypes.

The relationship between MLH1 and obesity remains unclear. Amongst the few published studies that have investigated this relationship, most have had a narrowed focus on the impact of lifestyle factors, such as diet, on MMR-deficient colon cancers. Campbell et al. (2009), for example, reported that a western diet pattern increased colon cancer risk in individuals with MLH1 polymorphisms[131]. Western diet patterns, characterized by increased consumption of high-fat and high-sugar foods, have previously been associated with increased risk for obesity and metabolic syndrome compared to traditional or Mediterranean diets[132, 133]. A randomized controlled trial reported an association of obesity and increased risk of colon cancer in patients with Lynch syndrome, a disorder that is caused by a defect in one or more of the MMR genes, typically MLH1, MSH2, and/or PMS2[118, 134]. Taking a more direct approach, Remely et al. (2017) reported a decreased expression of MLH1 and an increased level of CpG promoter methylation in mice fed high-fat diets compared to those on control diets[135]. Together, these data suggest underlying regulatory relationships between obesity and MLH1 status and prompt further investigations into how these two interact, particularly in the context of colon cancer development.

2.3.2. MutS Homologue 2 (MSH2)

MSH2 binds to either MSH6 or MSH3 to form the MutSα and MutSβ heterodimers, respectively, that can recognize mismatched or looped DNA and initiate the MMR mechanism. Like other proteins involved in MMR, defects in MSH2 can result in genomic instability and increased risk of several cancers. Similar to MLH1, MSH2 was found to be downregulated in aging hematopoietic stem cells[129]. Additionally, 12-month-old mice showed both increased hypermethylated MSH2 promoter regions and decreased MSH2 mRNA levels compared to 2-month-old mice, indicating a possible role for MSH2 in the aging process[136]. Further, mice deficient in MSH2 demonstrated increased mutation frequencies in spleen and colon tissues compared to their WT counterparts, similar to the observations seen in MLH1-deficient mice[125]. These results suggested that, like MLH1, deficiency in MSH2 may promote the aging process by an increased occurrence in microsatellite and genomic instability events. In support of this idea, Campbell et al. (2006) used MEFs from MSH2-deficient mice to study the possible roles of MMR, and specifically MSH2, on telomere maintenance and chromosome stability[137]. Indeed, a significant increase in chromosomal aberrations, including centrosome amplification, chromosome mis-segregation, and aneuploidy events were observed in MSH2-deficient MEFs compared to WT MEFs. Interestingly, MSH2 deficiency had only a slight impact on telomere capping defects, and no impact on telomere length or telomerase regulation. Thus, these results suggested that rather than having a direct effect on telomere maintenance, mutations in telomere sequences, and thus telomere function, may be due to the increased microsatellite instability events that occur due to defects in MMR.

As with MLH1, the relationship between MSH2 and obesity remains largely unknown. Most of the reported studies have focused on the impact of diet and gut microbiota on the development of CRC in MSH2-deficient systems. Dietary factors play a large role in the development of obesity and can lead to significant alterations in the microbiome and increased risk for the development of human diseases, as previously reviewed[138, 139]. Indeed, researchers found that the gut microbiome of mice genetically predisposed to CRC (i.e., MSH2-deficient mice with mutated APC) lead to polyp formation, hyperproliferation, and epithelial cell transformation, precursors for cancer development[140]. Interestingly, these events were significantly reduced when the mice were fed a low-carbohydrate diet, suggesting that the metabolites produced by the microbiome on diets reminiscent of western diet patterns were a driving force in the development of CRC[140]. While these experiments provide a solid foundation for understanding some of the mechanisms underlying the relationship between diet and cancer development, further studies are warranted to investigate the role(s) of MSH2 in obesity-related diseases and how these mechanisms can be utilized for the prevention and/or treatment of human disease.

2.4. Non-homologous End-joining (NHEJ)

Generated from both endogenous and exogenous sources, DSBs are among the most toxic and deleterious lesions that can occur in DNA. Mis-repair of these lesions can have severe consequences on the integrity of the genome and lead to a variety of mutations, including large deletions and chromosomal translocations[141]. Thus, it is imperative that DSBs are identified and repaired efficiently and in an error-free fashion. NHEJ is the primary mechanism for the repair of DSBs outside of S-phase. The NHEJ pathway and the proteins involved have been reviewed extensively[142-145]. Briefly, broken ends are recognized by the Ku heterodimer, composed of Ku70 and Ku80. Once bound, this heterodimer recruits most of the other components of the NHEJ pathway, including DNA-PKcs, X-ray cross-complementing protein 4 (XRCC4), DNA Ligase IV, XRCC4-like factor (XLF), and numerous other proteins required for generating compatible ends prior to ligation. Once compatible ends have been generated, XRCC4, DNA ligase IV, and XLF work together to seal the ends, effectively repairing the DSB (Figure 3). [142].

Often considered an error-prone process, studies have recently differentiated two primary mechanisms of repair with varying levels of fidelity; classical NHEJ (c-NHEJ), which is relatively error-free, and alternative NHEJ (A-NHEJ), which is generally more error-prone[146, 147]. Currently, it is unclear as to why one pathway is chosen over the other. During states of increased DNA damage or dysregulated DNA repair, such as with aging and/or obesity, it could be that the more mutagenic A-NHEJ pathway is able to compete with c-NHEJ and perpetuate the cycle of genetic instability. Several groups have observed that with age, the efficiency of NHEJ decreases, resulting in an accumulation of genomic instability[148-152]. Transgenic mice deficient in critical NHEJ proteins also demonstrated signs of premature aging, including proliferative defects, cellular senescence, and telomere shortening, all of which are prominent hallmarks of aging[153-156].

Reports demonstrating changes in the levels of NHEJ proteins with age are contradictory. Some groups have observed decreased Ku levels in human fibroblasts and lymphocytes with age and senescence[157, 158]. Other studies, however, reported that changes in Ku expression and activity were tissue-specific, with kidney, lung, testis, and liver displaying varying degrees of age-associated modulation[159]. In contrast, Li et al. (2016) analyzed the levels of several NHEJ proteins in 50 human eyelid fibroblasts from young and old doners and reported that XRCC4 and Ligase 4 proteins decreased with age, while Ku protein levels remained unchanged[152]. This observation of decreased XRCC4 expression with age has also been observed across several tissues in rats[160]. The cause for the discrepancies between these studies is unclear; however, it is evident that the modulation of NHEJ factors may be influenced by the aging process.

Interestingly, while both of the rodent studies mentioned above reported different proteins being affected by age, calorie restriction reduced the observed effects in both experiments, returning the expression and activity of the DNA repair factors to baseline levels[159, 160]. Ke et al. (2020) investigated this phenomenon further by placing NHEJ-reporter mice on short-term calorie restricted diets and analyzing NHEJ efficiency across multiple tissues[161]. Mice on calorie restricted diets had statistically significant increases in NHEJ repair efficiency in skin, lung, kidney, and brain, as well as increased protein levels of DNA-PK compared to mice on control diets. Although the diets were short-term, this study provided strong evidence that dietary intervention could aid in the prevention of an age-dependent decline in DNA repair and prompts further investigation into the impact of long-term calorie restriction on DNA repair efficiency.

Although there are few reported studies on calorie restriction and NHEJ, there are even fewer investigating the potential impact of obesity on NHEJ repair. Yu et al. (2018) reported that although mice on HFD demonstrated an accumulation of DSBs and apoptosis, NHEJ appeared unchanged[162]. It is important to note that this study used 53BP1 staining as an indicator of functional NHEJ instead of a transgenic-reporter mouse model. Given the effects of calorie restriction on NHEJ, and the accumulation of DSBs in mice on HFD, it would be worth investigating further for any potential relationships between obesity and NHEJ. The use of NHEJ-reporter mouse models or the development of functional NHEJ assays suitable for tissue extracts may provide a more unique and direct insight into the potential impact of obesity on NHEJ efficiency.

2.5. Homologous Recombination (HR)

In addition to NHEJ, HR is another primary mechanism for the repair of DSBs. However, unlike NHEJ, which can occur throughout the cell cycle, HR utilizes the sister chromatid as a template for error-free repair of broken DNA ends and thus only occurs during the S- and G2-phases of the cell cycle. Several recent publications have reviewed the current understanding of the complex and intricate processes underlying HR[163-166]. Briefly, the broken ends are recognized by a complex of proteins referred to as the MRN complex, including Meiotic recombination 11 (MRE11), Rad50, and Nijmegen breakage syndrome 1 (NBS1). Several additional proteins, including CtBP-interacting protein (CtIP), EXO1, Bloom Helicase-DNA replication ATP-dependent helicase/nuclease (BLM-DNA2) complex, and DNA repair protein Rad51 homologue 1 (Rad51) are then recruited for DNA end resection, strand invasion, and repair synthesis using the homologous sister chromatid as a template (Figure 3).

It is established that DNA damage, and in particular DSBs, increases with age. The extent to which this increase is due to an accumulation of DNA damage, decreased DNA repair, or a combination of both, remains unclear. Studies have demonstrated that, along with an increase in DSBs, the efficiency of HR repair decreases with age[152, 167]. Similar to NHEJ, there are conflicting reports in regard to the relative levels of HR repair proteins with age. Some groups have reported decreased levels of key HR proteins with age, including MRE11 and Rad51, in yeast and human lymphocytes[157, 168]. In contrast, Li et al. (2016) observed no changes in HR repair protein levels in aged human eyelid fibroblasts, and instead proposed that the impaired recruitment of Rad51 to DNA damage sites contributed to decreased HR efficiency[152]. Interestingly, in addition to decreased HR repair, another group utilizing Drosophila as a model of aging observed increased expression and recruitment of Rad51 to DNA damaged sites with age, opposite to the observations of the previously mentioned studies[167]. As overexpression of Rad51 had previously been reported to result in chromosomal aberrations and genetic instability, the authors proposed that excess Rad51 could potentially delay filament disassembly during repair synthesis and stall the repair of the break[169]. The contradictory reports in the levels of HR repair proteins during the aging process are likely attributed to both differences in the cell type and species utilized in the published studies.

Aside from alterations in protein levels, there is a growing body of evidence suggesting the involvement of the HR repair pathway in telomere maintenance. The MRN complex has been reported to negatively regulate access of TRF1 to telomere ends, allowing telomerase access to the telomere ends and promoting telomere lengthening[170]. Depletion of Rad51 and BRCA2 resulted in telomere attrition, fragility, and telomere-associated chromosomal abnormalities in various models, including fungus, MEFs, and human cells[171-173]. Rad51-depleted human cells showed increased 53BP1 foci formation and telomeric fragments after CRISPR-Cas9-induced DSBs, indicating decreased DNA repair and providing further evidence for the involvement of HR in telomeric DSB repair[174]. Mice deficient in Rad54, a Rad51-interacting protein during HR, demonstrated telomere shortening and chromosomal aberrations, suggesting a role in telomere protection[175]. Most recently, Feretzaki et al. (2020) reported that Rad51 facilitated the recruitment of TERRA, a long non-coding RNA reported to be involved in telomere maintenance, to telomere ends via R-loop formation[176, 177]. Together, it is evident that HR repair proteins play protective roles in preventing telomere fragility and instability, prominent hallmarks of aging.

While few in number, multiple reports have demonstrated increased DSBs, as assessed by γ-H2AX markers, during obesity[11, 162, 178, 179]. Moreover, Yu et al. (2018) reported evidence of decreased HR repair via fewer Rad51-positive foci in mice fed HFDs compared to control chow diets[162]. Although still requiring additional mechanistic evidence, these studies provide a foundation for future experiments to further investigate the relationship of obesity and DSB repair, and in particular how HR repair is impacted under conditions of obesity.

3. DNA Replication

DNA replication is among the most studied of the DNA metabolic processes and must occur before cell division can be completed. With trillions of cells dividing each day, it is imperative that the fidelity of DNA replication be preserved. Countless proteins are involved in ensuring accurate and efficient replication of the genome, and their dysregulation can result in disastrous consequences. Further, endogenous and exogenous factors can result in the stalling of the replication process. Unresolved replication stress can result in DNA damage, genomic instability, and cell death. In this section, we will discuss a few of the many proteins that are involved in maintaining DNA replication fidelity and their roles in aging and obesity, as well as how aging and obesity can modulate the functions of these proteins (see Table 1).

3.1. Werner Syndrome ATP-dependent Helicase (WRN)

A member of the RecQ family of helicases, WRN is a multifunctional protein involved in numerous DNA metabolic processes. Werner syndrome, the result of mutations in WRN, is characterized by premature aging and an increased risk of cancer[180, 181]. Additionally, patients with Werner syndrome are also at an increased risk for early onset T2DM, cataracts, and skin atrophy[180, 182]. Human cells depleted of WRN and primary cells from WRN patients displayed increased mitochondrial ROS, decreased NAD+ levels, and loss of mitochondrial morphology, indicating severe mitochondrial dysfunction[183]. Restoration of NAD+ to normal levels reversed these phenotypes and inhibited the accelerated aging observed in the cells[183]. Dysregulation of NAD+ levels and altered mitochondrial function are prominent in metabolic and premature aging disorders[184]. These findings suggested that mitochondrial dysfunction may be a significant contributor to the accelerated aging phenotypes observed in patients with Werner syndrome.

Helicase and exonuclease domains allow for the involvement of WRN in several DNA processes, including DNA replication, repair, and telomere maintenance[185], suggesting a primary function for WRN in maintaining genomic integrity. In support of this, primary fibroblasts from Werner syndrome donors demonstrated proliferative defects and early onset of cellular senescence compared to WT controls[186]. These results were recapitulated in fibroblasts depleted of WRN, showing a direct relationship between WRN function and the observed phenotypes[186]. Further, WRN was reported to be involved in the resolution of multiple alternative DNA structures, as depletion of WRN in human cells lead to increased DNA structure-induced genetic instability[187, 188]. There are many causes of early onset of cellular senescence, including DNA damage and telomeric dysfunction. Cells lacking WRN helicase activity showed defective telomere replication and deletions in telomeric regions[189]. Additional studies have reported roles for WRN in the repair of telomeric DNA damage, as cells depleted of WRN were more susceptible to telomeric DNA damage, telomere fusions, and telomere dysfunction[190-192]. Interestingly, a study by Shamanna et al. (2016) reported that WRN deficiency decreased NHEJ efficiency and promoted the more mutagenic A-NHEJ pathway, resulting in increased DNA end resections and telomere fusions in WRN-deficient MEFs[193]. These results suggested that WRN may play a role in determining pathway choice in DSB repair, and that increased A-NHEJ repair in WRN-deficient cells may be a contributing factor to the increased genomic instability and telomere dysfunction phenotypes observed in this progeroid disease.

Compared to WT control mice, WRN-deficient mice placed on HFD demonstrated increased weight gain, hyperinsulinemia, and insulin resistance, all signs of obesity and metabolic syndrome[194]. Although WT mice fed a HFD also showed similar phenotypes, they were not as severe as those observed in WRN-deficient mice. Another study reported by Massip et al. (2006) found that mice with a mutation in the helicase domain of the WRN protein had increased weight gain, hyperinsulinemia, triglyceride deposits, and cholesterol levels compared to WT mice[195]. The results from this study were particularly interesting, as they demonstrated that even in the absence of a HFD, mice with mutated WRN still developed clear signs of obesity and metabolic disorder. Interestingly, most patients with Werner syndrome have dysregulated insulin signaling, which is thought to be the main contributing factor for the early onset T2DM observed in these patients[194-196]. Further, the increase in insulin levels in mice deficient in WRN appeared to be slightly exacerbated with age, indicating that age-related alterations in metabolic signaling may also occur in the absence of functional WRN protein[197].

3.2. Bloom Syndrome (BLM) Protein

BLM is another RecQ helicase involved in maintaining genome stability and integrity. Deficiency in BLM leads to the developmental disorder known as Bloom syndrome, which is characterized by stunted growth, insulin resistance, immune deficiencies, and sensitivity to sunlight[198, 199]. As characteristics of Bloom syndrome are not consistent with other progeroid syndromes (e.g., graying hair, cataracts, skin atrophy, etc.), it is not currently classified as a progeroid disorder[200]. However, individuals with Bloom syndrome are far more susceptible to developing cancer, widely accepted as a disease of aging, and often die young due to health complications[198-200].

Multiple groups have demonstrated that knockdown of BLM in human primary fibroblast cells resulted in early onset of cellular senescence, a notable hallmark of aging and common phenotype of aging disorders[201, 202]. In vitro studies have demonstrated that BLM interacted with telomere-associated proteins such as TRF2, which stimulated its helicase activities[203]. BLM was also found to localize to a subset of telomeres in human cells, although telomere repeat lengths in Bloom syndrome cells were not significantly different than normal cells, suggesting its involvement may be more related to a surveillance mechanism than length maintenance[204]. However, a study by Barefield et al. (2012) demonstrated BLM association to telomeres in G2/M phases, and that lack of BLM resulted in chromosomal abnormalities, including chromosome fusions[205]. The association of BLM during late cell cycle phases suggests that BLM may be involved in post-replicative maintenance of telomeres. BLM has been previously reported to be involved in the HR repair pathway, which is only active during S/G2 phases when a sister chromatid is available, suggesting that the involvement of BLM in telomere homeostasis may be largely due to its involvement in DSB repair[206-208]. Further, BLM was reported to be involved in the resolution of alternative DNA structures as well as replication fork restart after replication stress, preventing DSB formation and preserving genome integrity[209, 210]. Together, these studies provide clear evidence that BLM is critical for maintaining genome stability.

Compared to WRN, there are fewer studies published on the roles of BLM in obesity and metabolic syndrome. In fact, most children with Bloom syndrome are often underweight and display a lack of appetite, although this is typically not carried over into adulthood with some adults developing obesity[199, 211]. Similar to WRN, many individuals with Bloom syndrome develop T2DM due to insulin resistance, although the mechanisms underlying the involvement of BLM in insulin signaling are currently unclear, and warrant further investigation[198].

3.3. Flap Endonuclease 1 (FEN1)

FEN1 is a multifunctional, structure-specific metallonuclease involved in several DNA metabolic processes, including lagging strand synthesis in DNA replication and long-patch BER. Its primary function is to resolve intermediate DNA structures to prevent genomic instability[17, 212]. Cells depleted of FEN1 often display triplet repeat instability, DNA-structure-induced mutagenesis, replication defects, and decreased cell survival[19, 212-214]. Further, mice with complete knockout of FEN1 are embryonically lethal and mice haploinsufficient for FEN1 are more susceptible to developing cancer, underscoring its critical roles in development and genome stability[215]. While there are limited published reports investigating FEN1 and obesity, a study published by Larsen et al. (2008) reported that mice with a mutated FEN1 nuclease domain developed obesity by 9 months of age compared to WT control mice[216]. The absence of alterations in insulin or glucose levels suggested that the development of obesity was independent of insulin resistance or diabetes. Currently, studies are lacking on the potential involvement of FEN1 in the development of obesity, or how the obese state could regulate FEN1 activity.

As mentioned throughout this review, prominent hallmarks of aging include telomere fragility and genomic instability, and many studies have demonstrated a role for FEN1 in these events. For example, a study by Saharia et al. (2008) demonstrated that depletion of FEN1 in human cells resulted in increased DNA damage at telomere sequences as well as telomere loss[217]. Further investigation revealed that telomere loss was observed in telomeres that underwent lagging strand synthesis, suggesting that telomere dysfunction in FEN1-depleted cells was due to replication stress. The WRN protein has been previously shown to associate with and stimulate the cleavage activity of FEN1 at telomeres[217, 218]. WRN has also been reported to be essential for replication fork restart, as well as forming a complex with FEN1 during replication fork stalling to process branch structures at the fork[219, 220]. Saharia et al. (2010) demonstrated that FEN1-depleted HeLa cells had less efficient replication fork restart after hydroxyurea treatment than WT control cells[221]. Therefore, it is likely that FEN1 is involved in the initiation of stalled replication forks during telomere replication along with WRN to prevent telomere dysfunction and instability.

Although its primary roles involve maintaining DNA replication integrity, FEN1 is also likely to be recruited to telomeres in response to DNA damage via roles in DNA repair pathways. Indeed, a study observing interactions of telomere binding proteins with DNA repair proteins revealed that TRF2, a telomere capping protein, associated with both FEN1 and Polβ[222]. Further, Sun et al. (2015) found that FEN1, along with Polβ and NTH1, was recruited to telomeres following site-specific DNA damage[223]. These data suggest a potential role for FEN1 in the resolution of DNA damage at telomeres via its involvement in the long-patch BER sub-pathway.

3.4. Twinkle Helicase

Twinkle is a mitochondrial DNA helicase critical for the replication of mitochondrial DNA. It functions primarily to unwind mitochondrial DNA at the replication fork and is thought to be the sole essential replicative helicase for mitochondrial DNA replication[224, 225]. Mutations in Twinkle result in mitochondrial diseases, such as progressive external ophthalmoplegia (PEO), mitochondrial DNA depletion syndromes, and Perrault syndrome[226-229]. Mitochondrial diseases are characterized by decreased mitochondrial respiration and mitochondrial dysfunction, common features observed in both aging and obese states. Some overlap exists in the phenotypes observed from Twinkle mutations and mutations in polymerase-γ (POLG). This is not unexpected, as Twinkle works in concert with POLG and the mitochondrial single-stranded DNA binding protein (mtSSB) to replicate mitochondrial DNA[226, 230]. Therefore, it is not surprising that mutations in key replisome factors would result in overlapping diseases, as they play crucial roles in the maintenance of mitochondrial DNA.

Along with a decline in mitochondrial respiration, aging mice also showed marked decreases in Twinkle protein levels and mitochondrial DNA copy number[231]. Additionally, a study analyzing mitochondrial DNA copy number in peripheral blood cells from over 1,000 patients demonstrated decreased copy number with age[232]. Mice overexpressing the Twinkle helicase have been reported to have increased mitochondrial DNA copy number[224]. Based on these findings, Foote et al. (2018) set out to determine if restoration of mitochondrial DNA copy number by overexpression of Twinkle could rescue the mitochondrial dysfunction observed in aging mice. Indeed, mice overexpressing Twinkle displayed significantly increased mitochondrial copy numbers and respiration rates compared to age-matched WT mice[231]. Further, mice overexpressing Twinkle appeared to have delayed vascular aging compared to age-matched WT mice, as demonstrated by a delay in the decline of arterial distensibility and arterial compliance. Together, these results highlight an interesting mechanism by which aging phenotypes such as mitochondrial dysfunction and vascular aging may be ameliorated by restoring mitochondrial DNA copy number and replication fidelity.

4. Concluding Remarks

DNA repair and replication proteins play crucial roles in the maintenance of genome integrity, and as such, deficiencies or dysregulation of these proteins can result in aberrant DNA repair, genetic instability, and the development of several diseases. Here, we have reviewed the existing literature investigating the roles and functions of various repair and replication proteins in aging and obesity, and how these proteins may be involved in the mechanisms underlying aging- and obesity-related disorders.

Some of these proteins, such as the OGG1 and NEIL proteins, have shown obvious roles for preventing common hallmarks of both aging and obesity, including the regulation of mitochondrial function, modulation of energy homeostasis, and prevention of genomic instability. Others, such as XP proteins, CS proteins, and the RecQ helicases WRN and BLM, have demonstrated roles primarily in the prevention of aging phenotypes, in which loss of these proteins leads to various progeroid syndromes. Nevertheless, the studies and findings discussed in this review provide clear and strong support for links between DNA repair fidelity, aging, and obesity and provide a solid foundation for further studies to investigate the underlying mechanisms involved.

Highlights.

  • Aging and obesity can lead to the dysregulation of DNA repair and replication proteins and/or pathways

  • Aging and obesity can result in increased DNA damage and decreased DNA repair, resulting in an accumulation of mutations and genomic instability

  • Deficiencies and/or the dysregulation of DNA repair and replication proteins can result in aberrant DNA repair, genomic instability, and the development of several diseases

  • Mutations in genes responsible for maintaining genome integrity can exacerbate aging- and obesity-related phenotypes and disorders

Acknowledgments

This work was supported by NIH/NCI grants to K.M.V (CA093729 and CA225029).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflicts of Interest

The authors have no conflicts of interest to declare.

References

  • [1].Huang JC, Svoboda DL, Reardon JT, Sancar A, Human nucleotide excision nuclease removes thymine dimers from DNA by incising the 22nd phosphodiester bond 5' and the 6th phosphodiester bond 3' to the photodimer, Proc Natl Acad Sci U S A, 89 (1992) 3664–3668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Krokan HE, Standal R, Slupphaug G, DNA glycosylases in the base excision repair of DNA, Biochem J, 325 ( Pt 1) (1997) 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Pannunzio NR, Watanabe G, Lieber MR, Nonhomologous DNA end-joining for repair of DNA double-strand breaks, J Biol Chem, 293 (2018) 10512–10523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Li X, Heyer WD, Homologous recombination in DNA repair and DNA damage tolerance, Cell Res, 18 (2008) 99–113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Lopez-Otin C, Blasco MA, Partridge L, Serrano M, Kroemer G, The hallmarks of aging, Cell, 153 (2013) 1194–1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Jin K, Modern Biological Theories of Aging, Aging Dis, 1 (2010) 72–74. [PMC free article] [PubMed] [Google Scholar]
  • [7].Bernadotte A, Mikhelson VM, Spivak IM, Markers of cellular senescence. Telomere shortening as a marker of cellular senescence, Aging (Albany NY), 8 (2016) 3–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Beckman KB, Ames BN, The free radical theory of aging matures, Physiol Rev, 78 (1998) 547–581. [DOI] [PubMed] [Google Scholar]
  • [9].Kompella P, Vasquez KM, Obesity and cancer: A mechanistic overview of metabolic changes in obesity that impact genetic instability, Mol Carcinog, 58 (2019) 1531–1550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Furukawa S, Fujita T, Shimabukuro M, Iwaki M, Yamada Y, Nakajima Y, Nakayama O, Makishima M, Matsuda M, Shimomura I, Increased oxidative stress in obesity and its impact on metabolic syndrome, J Clin Invest, 114 (2004) 1752–1761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Azzara A, Pirillo C, Giovannini C, Federico G, Scarpato R, Different repair kinetic of DSBs induced by mitomycin C in peripheral lymphocytes of obese and normal weight adolescents, Mutat Res, 789 (2016) 9–14. [DOI] [PubMed] [Google Scholar]
  • [12].Lindahl T, Nyberg B, Rate of depurination of native deoxyribonucleic acid, Biochemistry, 11 (1972) 3610–3618. [DOI] [PubMed] [Google Scholar]
  • [13].Dianov G, Lindahl T, Reconstitution of the DNA base excision-repair pathway, Curr Biol, 4 (1994) 1069–1076. [DOI] [PubMed] [Google Scholar]
  • [14].Kubota Y, Nash RA, Klungland A, Schar P, Barnes DE, Lindahl T, Reconstitution of DNA base excision-repair with purified human proteins: interaction between DNA polymerase beta and the XRCC1 protein, EMBO J, 15 (1996) 6662–6670. [PMC free article] [PubMed] [Google Scholar]
  • [15].Robertson AB, Klungland A, Rognes T, Leiros I, DNA repair in mammalian cells: Base excision repair: the long and short of it, Cell Mol Life Sci, 66 (2009) 981–993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Frosina G, Fortini P, Rossi O, Carrozzino F, Raspaglio G, Cox LS, Lane DP, Abbondandolo A, Dogliotti E, Two pathways for base excision repair in mammalian cells, J Biol Chem, 271 (1996) 9573–9578. [DOI] [PubMed] [Google Scholar]
  • [17].Prasad R, Dianov GL, Bohr VA, Wilson SH, FEN1 stimulation of DNA polymerase beta mediates an excision step in mammalian long patch base excision repair, J Biol Chem, 275 (2000) 4460–4466. [DOI] [PubMed] [Google Scholar]
  • [18].Starcevic D, Dalal S, Sweasy JB, Is there a link between DNA polymerase beta and cancer?, Cell Cycle, 3 (2004) 998–1001. [PubMed] [Google Scholar]
  • [19].Liu Y, Wilson SH, DNA base excision repair: a mechanism of trinucleotide repeat expansion, Trends Biochem Sci, 37 (2012) 162–172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Wallace SS, Base excision repair: a critical player in many games, DNA Repair (Amst), 19 (2014) 14–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Marsden CG, Dragon JA, Wallace SS, Sweasy JB, Base Excision Repair Variants in Cancer, Methods Enzymol, 591 (2017) 119–157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Xanthoudakis S, Smeyne RJ, Wallace JD, Curran T, The redox/DNA repair protein, Ref-1, is essential for early embryonic development in mice, Proc Natl Acad Sci U S A, 93 (1996) 8919–8923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [23].Gu H, Marth JD, Orban PC, Mossmann H, Rajewsky K, Deletion of a DNA polymerase beta gene segment in T cells using cell type-specific gene targeting, Science, 265 (1994) 103–106. [DOI] [PubMed] [Google Scholar]
  • [24].Sugo N, Aratani Y, Nagashima Y, Kubota Y, Koyama H, Neonatal lethality with abnormal neurogenesis in mice deficient in DNA polymerase beta, EMBO J, 19 (2000) 1397–1404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Shibutani S, Takeshita M, Grollman AP, Insertion of specific bases during DNA synthesis past the oxidation-damaged base 8-oxodG, Nature, 349 (1991) 431–434. [DOI] [PubMed] [Google Scholar]
  • [26].Akiyama M, Maki H, Sekiguchi M, Horiuchi T, A specific role of MutT protein: to prevent dG.dA mispairing in DNA replication, Proc Natl Acad Sci U S A, 86 (1989) 3949–3952. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Boiteux S, Radicella JP, Base excision repair of 8-hydroxyguanine protects DNA from endogenous oxidative stress, Biochimie, 81 (1999) 59–67. [DOI] [PubMed] [Google Scholar]
  • [28].de Souza-Pinto NC, Eide L, Hogue BA, Thybo T, Stevnsner T, Seeberg E, Klungland A, Bohr VA, Repair of 8-oxodeoxyguanosine lesions in mitochondrial dna depends on the oxoguanine dna glycosylase (OGG1) gene and 8-oxoguanine accumulates in the mitochondrial dna of OGG1-defective mice, Cancer Res, 61 (2001) 5378–5381. [PubMed] [Google Scholar]
  • [29].Grollman AP, Moriya M, Mutagenesis by 8-oxoguanine: an enemy within, Trends Genet, 9 (1993) 246–249. [DOI] [PubMed] [Google Scholar]
  • [30].Arai T, Kelly VP, Minowa O, Noda T, Nishimura S, High accumulation of oxidative DNA damage, 8-hydroxyguanine, in Mmh/Ogg1 deficient mice by chronic oxidative stress, Carcinogenesis, 23 (2002) 2005–2010. [DOI] [PubMed] [Google Scholar]
  • [31].Yuzefovych LV, Schuler AM, Chen J, Alvarez DF, Eide L, Ledoux SP, Wilson GL, Rachek LI, Alteration of mitochondrial function and insulin sensitivity in primary mouse skeletal muscle cells isolated from transgenic and knockout mice: role of ogg1, Endocrinology, 154 (2013) 2640–2649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [32].Sampath H, Vartanian V, Rollins MR, Sakumi K, Nakabeppu Y, Lloyd RS, 8-Oxoguanine DNA glycosylase (OGG1) deficiency increases susceptibility to obesity and metabolic dysfunction, PLoS One, 7 (2012) e51697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Scheffler K, Rachek L, You P, Rowe AD, Wang W, Kusnierczyk A, Kittelsen L, Bjoras M, Eide L, 8-oxoguanine DNA glycosylase (Ogg1) controls hepatic gluconeogenesis, DNA Repair (Amst), 61 (2018) 56–62. [DOI] [PubMed] [Google Scholar]
  • [34].Stuart JA, Bourque BM, de Souza-Pinto NC, Bohr VA, No evidence of mitochondrial respiratory dysfunction in OGG1-null mice deficient in removal of 8-oxodeoxyguanine from mitochondrial DNA, Free Radic Biol Med, 38 (2005) 737–745. [DOI] [PubMed] [Google Scholar]
  • [35].Bratic A, Larsson NG, The role of mitochondria in aging, J Clin Invest, 123 (2013) 951–957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Vartanian V, Tumova J, Dobrzyn P, Dobrzyn A, Nakabeppu Y, Lloyd RS, Sampath H, 8-oxoguanine DNA glycosylase (OGG1) deficiency elicits coordinated changes in lipid and mitochondrial metabolism in muscle, PLoS One, 12 (2017) e0181687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Daimon M, Oizumi T, Toriyama S, Karasawa S, Jimbu Y, Wada K, Kameda W, Susa S, Muramatsu M, Kubota I, Kawata S, Kato T, Association of the Ser326Cys polymorphism in the OGG1 gene with type 2 DM, Biochem Biophys Res Commun, 386 (2009) 26–29. [DOI] [PubMed] [Google Scholar]
  • [38].Thameem F, Puppala S, Lehman DM, Stern MP, Blangero J, Abboud HE, Duggirala R, Habib SL, The Ser(326)Cys polymorphism of 8-oxoguanine glycosylase 1 (OGG1) is associated with type 2 diabetes in Mexican Americans, Hum Hered, 70 (2010) 97–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Anderson SM, Naidoo RN, Ramkaran P, Asharam K, Muttoo S, Chuturgoon AA, OGG1 Ser326Cys polymorphism, HIV, obesity and air pollution exposure influences adverse birth outcome susceptibility, within South African Women, Reprod Toxicol, 79 (2018) 8–15. [DOI] [PubMed] [Google Scholar]
  • [40].Komakula SSB, Tumova J, Kumaraswamy D, Burchat N, Vartanian V, Ye H, Dobrzyn A, Lloyd RS, Sampath H, The DNA Repair Protein OGG1 Protects Against Obesity by Altering Mitochondrial Energetics in White Adipose Tissue, Sci Rep, 8 (2018) 14886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Harman D, Aging: a theory based on free radical and radiation chemistry, J Gerontol, 11 (1956) 298–300. [DOI] [PubMed] [Google Scholar]
  • [42].Chen JH, Ozanne SE, Hales CN, Methods of cellular senescence induction using oxidative stress, Methods Mol Biol, 371 (2007) 179–189. [DOI] [PubMed] [Google Scholar]
  • [43].Passos JF, Nelson G, Wang C, Richter T, Simillion C, Proctor CJ, Miwa S, Olijslagers S, Hallinan J, Wipat A, Saretzki G, Rudolph KL, Kirkwood TBL, von Zglinicki T, Feedback between p21 and reactive oxygen production is necessary for cell senescence, in: Mol Syst Biol, 2010, pp. 347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Chen Q, Fischer A, Reagan JD, Yan LJ, Ames BN, Oxidative DNA damage and senescence of human diploid fibroblast cells, Proc Natl Acad Sci U S A, 92 (1995) 4337–4341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Imam SZ, Karahalil B, Hogue BA, Souza-Pinto NC, Bohr VA, Mitochondrial and nuclear DNA-repair capacity of various brain regions in mouse is altered in an age-dependent manner, Neurobiol Aging, 27 (2006) 1129–1136. [DOI] [PubMed] [Google Scholar]
  • [46].Mikkelsen L, Bialkowski K, Risom L, Lohr M, Loft S, Moller P, Aging and defense against generation of 8-oxo-7,8-dihydro-2'-deoxyguanosine in DNA, Free Radic Biol Med, 47 (2009) 608–615. [DOI] [PubMed] [Google Scholar]
  • [47].Zhang Y, Zhang L, Zhang L, Bai J, Ge H, Liu P, Expression changes in DNA repair enzymes and mitochondrial DNA damage in aging rat lens, Mol Vis, 16 (2010) 1754–1763. [PMC free article] [PubMed] [Google Scholar]
  • [48].Picca A, Mankowski RT, Kamenov G, Anton SD, Manini TM, Buford TW, Saini SK, Calvani R, Landi F, Bernabei R, Marzetti E, Leeuwenburgh C, Advanced Age Is Associated with Iron Dyshomeostasis and Mitochondrial DNA Damage in Human Skeletal Muscle, in: Cells, 2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [49].Tosato M, Zamboni V, Ferrini A, Cesari M, The aging process and potential interventions to extend life expectancy, Clin Interv Aging, 2 (2007) 401–412. [PMC free article] [PubMed] [Google Scholar]
  • [50].Wang Z, Rhee DB, Lu J, Bohr CT, Zhou F, Vallabhaneni H, de Souza-Pinto NC, Liu Y, Characterization of oxidative guanine damage and repair in mammalian telomeres, PLoS Genet, 6 (2010) e1000951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Coluzzi E, Leone S, Sgura A, Oxidative Stress Induces Telomere Dysfunction and Senescence by Replication Fork Arrest, Cells, 8 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Rhee DB, Ghosh A, Lu J, Bohr VA, Liu Y, Factors that influence telomeric oxidative base damage and repair by DNA glycosylase OGG1, DNA Repair (Amst), 10 (2011) 34–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Oikawa S, Kawanishi S, Site-specific DNA damage at GGG sequence by oxidative stress may accelerate telomere shortening, FEBS Lett, 453 (1999) 365–368. [DOI] [PubMed] [Google Scholar]
  • [54].Singh A, Kukreti R, Saso L, Kukreti S, Oxidative Stress: Role and Response of Short Guanine Tracts at Genomic Locations, in: Int J Mol Sci, 2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Lu J, Liu Y, Deletion of Ogg1 DNA glycosylase results in telomere base damage and length alteration in yeast, EMBO J, 29 (2010) 398–409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [56].Zhou J, Fleming AM, Averill AM, Burrows CJ, Wallace SS, The NEIL glycosylases remove oxidized guanine lesions from telomeric and promoter quadruplex DNA structures, Nucleic Acids Res, 43 (2015) 4039–4054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Hazra TK, Kow YW, Hatahet Z, Imhoff B, Boldogh I, Mokkapati SK, Mitra S, Izumi T, Identification and characterization of a novel human DNA glycosylase for repair of cytosine-derived lesions, J Biol Chem, 277 (2002) 30417–30420. [DOI] [PubMed] [Google Scholar]
  • [58].Vartanian V, Lowell B, Minko IG, Wood TG, Ceci JD, George S, Ballinger SW, Corless CL, McCullough AK, Lloyd RS, The metabolic syndrome resulting from a knockout of the NEIL1 DNA glycosylase, Proc Natl Acad Sci U S A, 103 (2006) 1864–1869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Sampath H, Batra AK, Vartanian V, Carmical JR, Prusak D, King IB, Lowell B, Earley LF, Wood TG, Marks DL, McCullough AK, R. S. L, Variable penetrance of metabolic phenotypes and development of high-fat diet-induced adiposity in NEIL1-deficient mice, Am J Physiol Endocrinol Metab, 300 (2011) E724–734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [60].Roy LM, Jaruga P, Wood TG, McCullough AK, Dizdaroglu M, Lloyd RS, Human polymorphic variants of the NEIL1 DNA glycosylase, J Biol Chem, 282 (2007) 15790–15798. [DOI] [PubMed] [Google Scholar]
  • [61].Shinmura K, Tao H, Goto M, Igarashi H, Taniguchi T, Maekawa M, Takezaki T, Sugimura H, Inactivating mutations of the human base excision repair gene NEIL1 in gastric cancer, Carcinogenesis, 25 (2004) 2311–2317. [DOI] [PubMed] [Google Scholar]
  • [62].Park MH, Kwak SH, Kim KJ, Go MJ, Lee HJ, Kim KS, Hwang JY, Kimm K, Cho YM, Lee HK, Park KS, Lee JY, Identification of a genetic locus on chromosome 4q34-35 for type 2 diabetes with overweight, Exp Mol Med, 45 (2013) e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Skarpengland T, Holm S, Scheffler K, Gregersen I, Dahl TB, Suganthan R, Segers FM, Ostlie I, Otten JJ, Luna L, Ketelhuth DF, Lundberg AM, Neurauter CG, Hildrestrand G, Skjelland M, Bjorndal B, Svardal AM, Iversen PO, Hedin U, Nygard S, Olstad OK, Krohg-Sorensen K, Slupphaug G, Eide L, Kusnierczyk A, Folkersen L, Ueland T, Berge RK, Hansson GK, Biessen EA, Halvorsen B, Bjoras M, Aukrust P, Neil3-dependent base excision repair regulates lipid metabolism and prevents atherosclerosis in Apoe-deficient mice, Sci Rep, 6 (2016) 28337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Rolseth V, Runden-Pran E, Luna L, McMurray C, Bjoras M, Ottersen OP, Widespread distribution of DNA glycosylases removing oxidative DNA lesions in human and rodent brains, DNA Repair (Amst), 7 (2008) 1578–1588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Gredilla R, Garm C, Holm R, Bohr VA, Stevnsner T, Differential age-related changes in mitochondrial DNA repair activities in mouse brain regions, Neurobiol Aging, 31 (2010) 993–1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Canugovi C, Yoon JS, Feldman NH, Croteau DL, Mattson MP, Bohr VA, Endonuclease VIII-like 1 (NEIL1) promotes short-term spatial memory retention and protects from ischemic stroke-induced brain dysfunction and death in mice, Proc Natl Acad Sci U S A, 109 (2012) 14948–14953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [67].Zhou J, Chan J, Lambele M, Yusufzai T, Stumpff J, Opresko PL, Thali M, Wallace SS, NEIL3 Repairs Telomere Damage during S Phase to Secure Chromosome Segregation at Mitosis, Cell Rep, 20 (2017) 2044–2056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Zhou J, Liu M, Fleming AM, Burrows CJ, Wallace SS, Neil3 and NEIL1 DNA glycosylases remove oxidative damages from quadruplex DNA and exhibit preferences for lesions in the telomeric sequence context, J Biol Chem, 288 (2013) 27263–27272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Chakraborty A, Wakamiya M, Venkova-Canova T, Pandita RK, Aguilera-Aguirre L, Sarker AH, Singh DK, Hosoki K, Wood TG, Sharma G, Cardenas V, Sarkar PS, Sur S, Pandita TK, Boldogh I, Hazra TK, Neil2-null Mice Accumulate Oxidized DNA Bases in the Transcriptionally Active Sequences of the Genome and Are Susceptible to Innate Inflammation, J Biol Chem, 290 (2015) 24636–24648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Makasheva KA, Endutkin AV, Zharkov DO, Requirements for DNA bubble structure for efficient cleavage by helix-two-turn-helix DNA glycosylases, Mutagenesis, 35 (2020) 119–128. [DOI] [PubMed] [Google Scholar]
  • [71].Jayaraman L, Murthy KG, Zhu C, Curran T, Xanthoudakis S, Prives C, Identification of redox/repair protein Ref-1 as a potent activator of p53, Genes Dev, 11 (1997) 558–570. [DOI] [PubMed] [Google Scholar]
  • [72].Li M, Yang X, Lu X, Dai N, Zhang S, Cheng Y, Zhang L, Yang Y, Liu Y, Yang Z, Wang D, Wilson DM 3rd, APE1 deficiency promotes cellular senescence and premature aging features, Nucleic Acids Res, 46 (2018) 5664–5677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Madlener S, Strobel T, Vose S, Saydam O, Price BD, Demple B, Saydam N, Essential role for mammalian apurinic/apyrimidinic (AP) endonuclease Ape1/Ref-1 in telomere maintenance, Proc Natl Acad Sci U S A, 110 (2013) 17844–17849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Cabelof DC, Guo Z, Raffoul JJ, Sobol RW, Wilson SH, Richardson A, Heydari AR, Base excision repair deficiency caused by polymerase beta haploinsufficiency: accelerated DNA damage and increased mutational response to carcinogens, Cancer Res, 63 (2003) 5799–5807. [PubMed] [Google Scholar]
  • [75].Cabelof DC, Ikeno Y, Nyska A, Busuttil RA, Anyangwe N, Vijg J, Matherly LH, Tucker JD, Wilson SH, Richardson A, Heydari AR, Haploinsufficiency in DNA polymerase beta increases cancer risk with age and alters mortality rate, Cancer Res, 66 (2006) 7460–7465. [DOI] [PubMed] [Google Scholar]
  • [76].Ahmed AA, Smoczer C, Pace B, Patterson D, Cress Cabelof D, Loss of DNA polymerase beta induces cellular senescence, Environ Mol Mutagen, 59 (2018) 603–612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Hosono S, Matsuo K, Ito H, Oze I, Hirose K, Watanabe M, Nakanishi T, Tajima K, Tanaka H, Polymorphisms in base excision repair genes are associated with endometrial cancer risk among postmenopausal Japanese women, Int J Gynecol Cancer, 23 (2013) 1561–1568. [DOI] [PubMed] [Google Scholar]
  • [78].Yesil-Devecioglu T, Dayan A, Demirtunc R, Sardas S, Role of DNA repair genes XRCC3 and XRCC1 in predisposition to type 2 diabetes mellitus and diabetic nephropathy, Endocrinol Diabetes Nutr, 66 (2019) 90–98. [DOI] [PubMed] [Google Scholar]
  • [79].Xu C, Xu J, Ji G, Liu Q, Shao W, Chen Y, Gu J, Weng Z, Zhang X, Wang Y, Gu A, Deficiency of X-ray repair cross-complementing group 1 in primordial germ cells contributes to male infertility, FASEB J, 33 (2019) 7427–7436. [DOI] [PubMed] [Google Scholar]
  • [80].Tebbs RS, Flannery ML, Meneses JJ, Hartmann A, Tucker JD, Thompson LH, Cleaver JE, Pedersen RA, Requirement for the Xrcc1 DNA base excision repair gene during early mouse development, Dev Biol, 208 (1999) 513–529. [DOI] [PubMed] [Google Scholar]
  • [81].McNeill DR, Lin PC, Miller MG, Pistell PJ, de Souza-Pinto NC, Fishbein KW, Spencer RG, Liu Y, Pettan-Brewer C, Ladiges WC, Wilson DM 3rd, XRCC1 haploinsufficiency in mice has little effect on aging, but adversely modifies exposure-dependent susceptibility, Nucleic Acids Res, 39 (2011) 7992–8004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Araujo SJ, Tirode F, Coin F, Pospiech H, Syvaoja JE, Stucki M, Hubscher U, Egly JM, Wood RD, Nucleotide excision repair of DNA with recombinant human proteins: definition of the minimal set of factors, active forms of TFIIH, and modulation by CAK, Genes Dev, 14 (2000) 349–359. [PMC free article] [PubMed] [Google Scholar]
  • [83].Faridounnia M, Folkers GE, Boelens R, Function and Interactions of ERCC1-XPF in DNA Damage Response, in: Molecules, 2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Mori T, Yousefzadeh MJ, Faridounnia M, Chong JX, Hisama FM, Hudgins L, Mercado G, Wade EA, Barghouthy AS, Lee L, Martin GM, Nickerson DA, Bamshad MJ, G. University of Washington Center for Mendelian, Niedernhofer LJ, Oshima J, ERCC4 variants identified in a cohort of patients with segmental progeroid syndromes, Hum Mutat, 39 (2018) 255–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Kim DE, Dolle MET, Vermeij WP, Gyenis A, Vogel K, Hoeijmakers JHJ, Wiley CD, Davalos AR, Hasty P, Desprez PY, Campisi J, Deficiency in the DNA repair protein ERCC1 triggers a link between senescence and apoptosis in human fibroblasts and mouse skin, Aging Cell, 19 (2020) e13072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Niedernhofer LJ, Garinis GA, Raams A, Lalai AS, Robinson AR, Appeldoorn E, Odijk H, Oostendorp R, Ahmad A, van Leeuwen W, Theil AF, Vermeulen W, van der Horst GT, Meinecke P, Kleijer WJ, Vijg J, Jaspers NG, Hoeijmakers JH, A new progeroid syndrome reveals that genotoxic stress suppresses the somatotroph axis, Nature, 444 (2006) 1038–1043. [DOI] [PubMed] [Google Scholar]
  • [87].Jaspers NGJ, Raams A, Silengo MC, Wijgers N, Niedernhofer LJ, Robinson AR, Giglia-Mari G, Hoogstraten D, Kleijer WJ, Hoeijmakers JHJ, Vermeulen W, First Reported Patient with Human ERCC1 Deficiency Has Cerebro-Oculo-Facio-Skeletal Syndrome with a Mild Defect in Nucleotide Excision Repair and Severe Developmental Failure, in: Am J Hum Genet, 2007, pp. 457–466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Kashiyama K, Nakazawa Y, Pilz DT, Guo C, Shimada M, Sasaki K, Fawcett H, Wing JF, Lewin SO, Carr L, Li TS, Yoshiura K, Utani A, Hirano A, Yamashita S, Greenblatt D, Nardo T, Stefanini M, McGibbon D, Sarkany R, Fassihi H, Takahashi Y, Nagayama Y, Mitsutake N, Lehmann AR, Ogi T, Malfunction of nuclease ERCC1-XPF results in diverse clinical manifestations and causes Cockayne syndrome, xeroderma pigmentosum, and Fanconi anemia, Am J Hum Genet, 92 (2013) 807–819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].McWhir J, Selfridge J, Harrison DJ, Squires S, Melton DW, Mice with DNA repair gene (ERCC-1) deficiency have elevated levels of p53, liver nuclear abnormalities and die before weaning, Nat Genet, 5 (1993) 217–224. [DOI] [PubMed] [Google Scholar]
  • [90].Chen Q, Liu K, Robinson AR, Clauson CL, Blair HC, Robbins PD, Niedernhofer LJ, Ouyang H, DNA damage drives accelerated bone aging via an NF-kappaB-dependent mechanism, J Bone Miner Res, 28 (2013) 1214–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Yousefzadeh MJ, Zhao J, Bukata C, Wade EA, McGowan SJ, Angelini LA, Bank MP, Gurkar AU, McGuckian CA, Calubag MF, Kato JI, Burd CE, Robbins PD, Niedernhofer LJ, Tissue specificity of senescent cell accumulation during physiologic and accelerated aging of mice, Aging Cell, 19 (2020) e13094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Zhu XD, Niedernhofer L, Kuster B, Mann M, Hoeijmakers JH, de Lange T, ERCC1/XPF removes the 3' overhang from uncapped telomeres and represses formation of telomeric DNA-containing double minute chromosomes, Mol Cell, 12 (2003) 1489–1498. [DOI] [PubMed] [Google Scholar]
  • [93].Wu Y, Mitchell TR, Zhu XD, Human XPF controls TRF2 and telomere length maintenance through distinctive mechanisms, Mech Ageing Dev, 129 (2008) 602–610. [DOI] [PubMed] [Google Scholar]
  • [94].Robinson AR, Yousefzadeh MJ, Rozgaja TA, Wang J, Li X, Tilstra JS, Feldman CH, Gregg SQ, Johnson CH, Skoda EM, Frantz MC, Bell-Temin H, Pope-Varsalona H, Gurkar AU, Nasto LA, Robinson RAS, Fuhrmann-Stroissnigg H, Czerwinska J, McGowan SJ, Cantu-Medellin N, Harris JB, Maniar S, Ross MA, Trussoni CE, LaRusso NF, Cifuentes-Pagano E, Pagano PJ, Tudek B, Vo NV, Rigatti LH, Opresko PL, Stolz DB, Watkins SC, Burd CE, Croix CMS, Siuzdak G, Yates NA, Robbins PD, Wang Y, Wipf P, Kelley EE, Niedernhofer LJ, Spontaneous DNA damage to the nuclear genome promotes senescence, redox imbalance and aging, Redox Biol, 17 (2018) 259–273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Czerwinska J, Nowak M, Wojtczak P, Dziuban-Lech D, Ciesla JM, Kolata D, Gajewska B, Baranczyk-Kuzma A, Robinson AR, Shane HL, Gregg SQ, Rigatti LH, Yousefzadeh MJ, Gurkar AU, McGowan SJ, Kosicki K, Bednarek M, Zarakowska E, Gackowski D, Olinski R, Speina E, Niedernhofer LJ, Tudek B, ERCC1-deficient cells and mice are hypersensitive to lipid peroxidation, Free Radic Biol Med, 124 (2018) 79–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Karakasilioti I, Kamileri I, Chatzinikolaou G, Kosteas T, Vergadi E, Robinson AR, Tsamardinos I, Rozgaja TA, Siakouli S, Tsatsanis C, Niedernhofer LJ, Garinis GA, DNA damage triggers a chronic autoinflammatory response, leading to fat depletion in NER progeria, Cell Metab, 18 (2013) 403–415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Bowden NA, Beveridge NJ, Ashton KA, Baines KJ, Scott RJ, Understanding Xeroderma Pigmentosum Complementation Groups Using Gene Expression Profiling after UV-Light Exposure, Int J Mol Sci, 16 (2015) 15985–15996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [98].Masutani C, Araki M, Yamada A, Kusumoto R, Nogimori T, Maekawa T, Iwai S, Hanaoka F, Xeroderma pigmentosum variant (XP-V) correcting protein from HeLa cells has a thymine dimer bypass DNA polymerase activity, EMBO J, 18 (1999) 3491–3501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Feltes BC, Bonatto D, Overview of xeroderma pigmentosum proteins architecture, mutations and post-translational modifications, Mutat Res Rev Mutat Res, 763 (2015) 306–320. [DOI] [PubMed] [Google Scholar]
  • [100].Fang EF, Scheibye-Knudsen M, Brace LE, Kassahun H, SenGupta T, Nilsen H, Mitchell JR, Croteau DL, Bohr VA, Defective mitophagy in XPA via PARP-1 hyperactivation and NAD(+)/SIRT1 reduction, Cell, 157 (2014) 882–896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Stout GJ, Blasco MA, Telomere length and telomerase activity impact the UV sensitivity syndrome xeroderma pigmentosum C, Cancer Res, 73 (2013) 1844–1854. [DOI] [PubMed] [Google Scholar]
  • [102].Harada YN, Shiomi N, Koike M, Ikawa M, Okabe M, Hirota S, Kitamura Y, Kitagawa M, Matsunaga T, Nikaido O, Shiomi T, Postnatal growth failure, short life span, and early onset of cellular senescence and subsequent immortalization in mice lacking the xeroderma pigmentosum group G gene, Mol Cell Biol, 19 (1999) 2366–2372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [103].Gopalakrishnan K, Low GKM, Ting APL, Srikanth P, Slijepcevic P, Hande MP, Hydrogen peroxide induced genomic instability in nucleotide excision repair-deficient lymphoblastoid cells, Genome Integr, 1 (2010) 16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Wu M, Audet A, Cusic J, Seeger D, Cochran R, Ghribi O, Broad DNA repair responses in neural injury are associated with activation of the IL-6 pathway in cholesterol-fed rabbits, J Neurochem, 111 (2009) 1011–1021. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • [105].Chen YW, Harris RA, Hatahet Z, Chou KM, Ablation of XP-V gene causes adipose tissue senescence and metabolic abnormalities, Proc Natl Acad Sci U S A, 112 (2015) E4556–4564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [106].Gorgoulis V, Adams PD, Alimonti A, Bennett DC, Bischof O, Bishop C, Campisi J, Collado M, Evangelou K, Ferbeyre G, Gil J, Hara E, Krizhanovsky V, Jurk D, Maier AB, Narita M, Niedernhofer L, Passos JF, Robbins PD, Schmitt CA, Sedivy J, Vougas K, von Zglinicki T, Zhou D, Serrano M, Demaria M, Cellular Senescence: Defining a Path Forward, Cell, 179 (2019) 813–827. [DOI] [PubMed] [Google Scholar]
  • [107].Vessoni AT, Guerra CCC, Kajitani GS, Nascimento LLS, Garcia CCM, Cockayne Syndrome: The many challenges and approaches to understand a multifaceted disease, Genet Mol Biol, 43 (2020) e20190085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [108].Cockayne EA, Dwarfism with retinal atrophy and deafness, Arch Dis Child, 11 (1936) 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].van der Weegen Y, Golan-Berman H, Mevissen TET, Apelt K, González-Prieto R, Goedhart J, Heilbrun EE, Vertegaal ACO, van den Heuvel D, Walter JC, Adar S, Luijsterburg MS, The cooperative action of CSB, CSA, and UVSSA target TFIIH to DNA damage-stalled RNA polymerase II, in: Nat Commun, 2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Cordisco S, Tinaburri L, Teson M, Orioli D, Cardin R, Degan P, Stefanini M, Zambruno G, Guerra L, Dellambra E, Cockayne Syndrome Type A Protein Protects Primary Human Keratinocytes from Senescence, J Invest Dermatol, 139 (2019) 38–50. [DOI] [PubMed] [Google Scholar]
  • [111].Crochemore C, Fernández-Molina C, Montagne B, Salles A, Ricchetti M, CSB promoter downregulation via histone H3 hypoacetylation is an early determinant of replicative senescence, in: Nat Commun, 2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Lee JH, Demarest TG, Babbar M, Kim EW, Okur MN, De S, Croteau DL, Bohr VA, Cockayne syndrome group B deficiency reduces H3K9me3 chromatin remodeler SETDB1 and exacerbates cellular aging, in: Nucleic Acids Res, 2019, pp. 8548–8562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Tan J, Duan M, Yadav T, Phoon L, Wang X, Zhang JM, Zou L, Lan L, An R-loop-initiated CSB-RAD52-POLD3 pathway suppresses ROS-induced telomeric DNA breaks, Nucleic Acids Res, 48 (2020) 1285–1300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Feng E, Batenburg NL, Walker JR, Ho A, Mitchell TRH, Qin J, Zhu XD, CSB cooperates with SMARCAL1 to maintain telomere stability in ALT cells, J Cell Sci, 133 (2020). [DOI] [PubMed] [Google Scholar]
  • [115].Batenburg NL, Mitchell TR, Leach DM, Rainbow AJ, Zhu XD, Cockayne Syndrome group B protein interacts with TRF2 and regulates telomere length and stability, Nucleic Acids Res, 40 (2012) 9661–9674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Li GM, Mechanisms and functions of DNA mismatch repair, Cell Res, 18 (2008) 85–98. [DOI] [PubMed] [Google Scholar]
  • [117].Hsieh P, Yamane K, DNA mismatch repair: molecular mechanism, cancer, and ageing, Mech Ageing Dev, 129 (2008) 391–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Martin-Lopez JV, Fishel R, The mechanism of mismatch repair and the functional analysis of mismatch repair defects in Lynch syndrome, Fam Cancer, 12 (2013) 159–168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Loeb KR, Loeb LA, Genetic Instability and the Mutator Phenotype : Studies in Ulcerative Colitis, in: Am J Pathol, 1999, pp. 1621–1626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Loeb LA, Microsatellite instability: marker of a mutator phenotype in cancer, Cancer Res, 54 (1994) 5059–5063. [PubMed] [Google Scholar]
  • [121].Poynter JN, Siegmund KD, Weisenberger DJ, Long TI, Thibodeau SN, Lindor N, Young J, Jenkins MA, Hopper JL, Baron JA, Buchanan D, Casey G, Levine AJ, Le Marchand L, Gallinger S, Bapat B, Potter JD, Newcomb PA, Haile RW, Laird PW, I. Colon Cancer Family Registry, Molecular characterization of MSI-H colorectal cancer by MLHI promoter methylation, immunohistochemistry, and mismatch repair germline mutation screening, Cancer Epidemiol Biomarkers Prev, 17 (2008) 3208–3215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Campbell PT, Jacobs ET, Ulrich CM, Figueiredo JC, Poynter JN, McLaughlin JR, Haile RW, Jacobs EJ, Newcomb PA, Potter JD, Le Marchand L, Green RC, Parfrey P, Younghusband HB, Cotterchio M, Gallinger S, Jenkins MA, Hopper JL, Baron JA, Thibodeau SN, Lindor NM, Limburg PJ, Martinez ME, Colon Cancer Family R, Case-control study of overweight, obesity, and colorectal cancer risk, overall and by tumor microsatellite instability status, J Natl Cancer Inst, 102 (2010) 391–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Jiricny J, Nystrom-Lahti M, Mismatch repair defects in cancer, Curr Opin Genet Dev, 10 (2000) 157–161. [DOI] [PubMed] [Google Scholar]
  • [124].Edelmann W, Cohen PE, Kane M, Lau K, Morrow B, Bennett S, Umar A, Kunkel T, Cattoretti G, Chaganti R, Pollard JW, Kolodner RD, Kucherlapati R, Meiotic pachytene arrest in MLH1-deficient mice, Cell, 85 (1996) 1125–1134. [DOI] [PubMed] [Google Scholar]
  • [125].Hegan DC, Narayanan L, Jirik FR, Edelmann W, Liskay RM, Glazer PM, Differing patterns of genetic instability in mice deficient in the mismatch repair genes Pms2, Mlh1, Msh2, Msh3 and Msh6, Carcinogenesis, 27 (2006) 2402–2408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Kenyon J, Fu P, Lingas K, Thomas E, Saurastri A, Santos Guasch G, Wald D, Gerson SL, Humans accumulate microsatellite instability with acquired loss of MLH1 protein in hematopoietic stem and progenitor cells as a function of age, Blood, 120 (2012) 3229–3236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Kim DJ, Yi SM, Lee SY, Kang HS, Choi YH, Song YW, Park SC, Association between the MLH1 gene and longevity, Hum Genet, 119 (2006) 353–354. [DOI] [PubMed] [Google Scholar]
  • [128].Arai T, Kasahara I, Sawabe M, Honma N, Aida J, Tabubo K, Role of methylation of the hMLH1 gene promoter in the development of gastric and colorectal carcinoma in the elderly, Geriatr Gerontol Int, 10 Suppl 1 (2010) S207–212. [DOI] [PubMed] [Google Scholar]
  • [129].Lian X, Dong Y, Zhao M, Liang Y, Jiang W, Li W, Zhang L, RNA-Seq analysis of differentially expressed genes relevant to mismatch repair in aging hematopoietic stem-progenitor cells, J Cell Biochem, (2019). [DOI] [PubMed] [Google Scholar]
  • [130].Edelmann W, Yang K, Kuraguchi M, Heyer J, Lia M, Kneitz B, Fan K, Brown AM, Lipkin M, Kucherlapati R, Tumorigenesis in Mlh1 and Mlh1/Apc1638N mutant mice, Cancer Res, 59 (1999) 1301–1307. [PubMed] [Google Scholar]
  • [131].Campbell PT, Curtin K, Ulrich CM, Samowitz WS, Bigler J, Velicer CM, Caan B, Potter JD, Slattery ML, Mismatch repair polymorphisms and risk of colon cancer, tumour microsatellite instability and interactions with lifestyle factors, Gut, 58 (2009) 661–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [132].Marwitz SE, Woodie LN, Blythe SN, Western-style diet induces insulin insensitivity and hyperactivity in adolescent male rats, Physiol Behav, 151 (2015) 147–154. [DOI] [PubMed] [Google Scholar]
  • [133].Shively CA, Appt SE, Vitolins MZ, Uberseder B, Michalson KT, Silverstein-Metzler MG, Register TC, Mediterranean versus Western Diet Effects on Caloric Intake, Obesity, Metabolism, and Hepatosteatosis in Nonhuman Primates, Obesity (Silver Spring), 27 (2019) 777–784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Movahedi M, Bishop DT, Macrae F, Mecklin JP, Moeslein G, Olschwang S, Eccles D, Evans DG, Maher ER, Bertario L, Bisgaard ML, Dunlop MG, Ho JW, Hodgson SV, Lindblom A, Lubinski J, Morrison PJ, Murday V, Ramesar RS, Side L, Scott RJ, Thomas HJ, Vasen HF, Burn J, Mathers JC, Obesity, Aspirin, and Risk of Colorectal Cancer in Carriers of Hereditary Colorectal Cancer: A Prospective Investigation in the CAPP2 Study, J Clin Oncol, 33 (2015) 3591–3597. [DOI] [PubMed] [Google Scholar]
  • [135].Remely M, Ferk F, Sterneder S, Setayesh T, Kepcija T, Roth S, Noorizadeh R, Greunz M, Rebhan I, Wagner KH, Knasmuller S, Haslberger A, Vitamin E Modifies High-Fat Diet-Induced Increase of DNA Strand Breaks, and Changes in Expression and DNA Methylation of Dnmt1 and MLH1 in C57BL/6J Male Mice, Nutrients, 9 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [136].Conde-Perezprina JC, Luna-Lopez A, Lopez-Diazguerrero NE, Damian-Matsumura P, Zentella A, Konigsberg M, Msh2 promoter region hypermethylation as a marker of aging-related deterioration in old retired female breeder mice, Biogerontology, 9 (2008) 325–334. [DOI] [PubMed] [Google Scholar]
  • [137].Campbell MR, Wang Y, Andrew SE, Liu Y, Msh2 deficiency leads to chromosomal abnormalities, centrosome amplification, and telomere capping defect, Oncogene, 25 (2006) 2531–2536. [DOI] [PubMed] [Google Scholar]
  • [138].Sivamaruthi BS, Kesika P, Suganthy N, Chaiyasut C, A Review on Role of Microbiome in Obesity and Antiobesity Properties of Probiotic Supplements, Biomed Res Int, 2019 (2019) 3291367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [139].Bouter KE, van Raalte DH, Groen AK, Nieuwdorp M, Role of the Gut Microbiome in the Pathogenesis of Obesity and Obesity-Related Metabolic Dysfunction, Gastroenterology, 152 (2017) 1671–1678. [DOI] [PubMed] [Google Scholar]
  • [140].Belcheva A, Irrazabal T, Robertson SJ, Streutker C, Maughan H, Rubino S, Moriyama EH, Copeland JK, Surendra A, Kumar S, Green B, Geddes K, Pezo RC, Navarre WW, Milosevic M, Wilson BC, Girardin SE, Wolever TMS, Edelmann W, Guttman DS, Philpott DJ, Martin A, Gut microbial metabolism drives transformation of MSH2-deficient colon epithelial cells, Cell, 158 (2014) 288–299. [DOI] [PubMed] [Google Scholar]
  • [141].Iliakis G, Murmann T, Soni A, Alternative end-joining repair pathways are the ultimate backup for abrogated classical non-homologous end-joining and homologous recombination repair: Implications for the formation of chromosome translocations, Mutat Res Genet Toxicol Environ Mutagen, 793 (2015) 166–175. [DOI] [PubMed] [Google Scholar]
  • [142].Davis AJ, Chen DJ, DNA double strand break repair via non-homologous end-joining, Transl Cancer Res, 2 (2013) 130–143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].Mao Z, Bozzella M, Seluanov A, Gorbunova V, DNA repair by nonhomologous end joining and homologous recombination during cell cycle in human cells, Cell Cycle, 7 (2008) 2902–2906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Lieber MR, The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway, Annu Rev Biochem, 79 (2010) 181–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [145].Davis AJ, Chen BP, Chen DJ, DNA-PK: a dynamic enzyme in a versatile DSB repair pathway, DNA Repair (Amst), 17 (2014) 21–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [146].Chakraborty A, Tapryal N, Venkova T, Horikoshi N, Pandita RK, Sarker AH, Sarkar PS, Pandita TK, Hazra TK, Classical non-homologous end-joining pathway utilizes nascent RNA for error-free double-strand break repair of transcribed genes, Nat Commun, 7 (2016) 13049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [147].Betermier M, Bertrand P, Lopez BS, Is non-homologous end-joining really an inherently error-prone process?, PLoS Genet, 10 (2014) e1004086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [148].Seluanov A, Mittelman D, Pereira-Smith OM, Wilson JH, Gorbunova V, DNA end joining becomes less efficient and more error-prone during cellular senescence, Proc Natl Acad Sci U S A, 101 (2004) 7624–7629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [149].Vyjayanti VN, Rao KS, DNA double strand break repair in brain: reduced NHEJ activity in aging rat neurons, Neurosci Lett, 393 (2006) 18–22. [DOI] [PubMed] [Google Scholar]
  • [150].Kalfalah F, Seggewiss S, Walter R, Tigges J, Moreno-Villanueva M, Burkle A, Ohse S, Busch H, Boerries M, Hildebrandt B, Royer-Pokora B, Boege F, Structural chromosome abnormalities, increased DNA strand breaks and DNA strand break repair deficiency in dermal fibroblasts from old female human donors, Aging (Albany NY), 7 (2015) 110–122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Mayer PJ, Lange CS, Bradley MO, Nichols WW, Age-dependent decline in rejoining of X-ray-induced DNA double-strand breaks in normal human lymphocytes, Mutat Res, 219 (1989) 95–100. [DOI] [PubMed] [Google Scholar]
  • [152].Li Z, Zhang W, Chen Y, Guo W, Zhang J, Tang H, Xu Z, Zhang H, Tao Y, Wang F, Jiang Y, Sun FL, Mao Z, Impaired DNA double-strand break repair contributes to the age-associated rise of genomic instability in humans, Cell Death Differ, 23 (2016) 1765–1777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Holcomb VB, Vogel H, Hasty P, Deletion of Ku80 causes early aging independent of chronic inflammation and Rag-1-induced DSBs, Mech Ageing Dev, 128 (2007) 601–608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [154].Espejel S, Klatt P, Menissier-de Murcia J, Martin-Caballero J, Flores JM, Taccioli G, de Murcia G, Blasco MA, Impact of telomerase ablation on organismal viability, aging, and tumorigenesis in mice lacking the DNA repair proteins PARP-1, Ku86, or DNA-PKcs, J Cell Biol, 167 (2004) 627–638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Li H, Vogel H, Holcomb VB, Gu Y, Hasty P, Deletion of Ku70, Ku80, or both causes early aging without substantially increased cancer, Mol Cell Biol, 27 (2007) 8205–8214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Vaidya A, Mao Z, Tian X, Spencer B, Seluanov A, Gorbunova V, Knock-in reporter mice demonstrate that DNA repair by non-homologous end joining declines with age, PLoS Genet, 10 (2014) e1004511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [157].Ju YJ, Lee KH, Park JE, Yi YS, Yun MY, Ham YH, Kim TJ, Choi HM, Han GJ, Lee JH, Lee J, Han JS, Lee KM, Park GH, Decreased expression of DNA repair proteins Ku70 and Mre11 is associated with aging and may contribute to the cellular senescence, Exp Mol Med, 38 (2006) 686–693. [DOI] [PubMed] [Google Scholar]
  • [158].Seluanov A, Danek J, Hause N, Gorbunova V, Changes in the level and distribution of Ku proteins during cellular senescence, DNA Repair (Amst), 6 (2007) 1740–1748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Um JH, Kim SJ, Kim DW, Ha MY, Jang JH, Kim DW, Chung BS, Kang CD, Kim SH, Tissue-specific changes of DNA repair protein Ku and mtHSP70 in aging rats and their retardation by caloric restriction, Mech Ageing Dev, 124 (2003) 967–975. [DOI] [PubMed] [Google Scholar]
  • [160].Lee JE, Heo JI, Park SH, Kim JH, Kho YJ, Kang HJ, Chung HY, Yoon JL, Lee JY, Calorie restriction (CR) reduces age-dependent decline of non-homologous end joining (NHEJ) activity in rat tissues, Exp Gerontol, 46 (2011) 891–896. [DOI] [PubMed] [Google Scholar]
  • [161].Ke Z, Firsanov D, Spencer B, Seluanov A, Gorbunova V, Short-term calorie restriction enhances DNA repair by non-homologous end joining in mice, NPJ Aging Mech Dis, 6 (2020) 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [162].Yu H, Harrison FE, Xia F, Altered DNA repair; an early pathogenic pathway in Alzheimer's disease and obesity, Sci Rep, 8 (2018) 5600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [163].Ayra-Plasencia J, Machin F, DNA double-strand breaks in telophase lead to coalescence between segregated sister chromatid loci, Nat Commun, 10 (2019) 2862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [164].Sun Y, McCorvie TJ, Yates LA, Zhang X, Structural basis of homologous recombination, Cell Mol Life Sci, 77 (2020) 3–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [165].Wright WD, Shah SS, Heyer WD, Homologous recombination and the repair of DNA double-strand breaks, J Biol Chem, 293 (2018) 10524–10535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [166].Jasin M, Rothstein R, Repair of strand breaks by homologous recombination, Cold Spring Harb Perspect Biol, 5 (2013) a012740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [167].Delabaere L, Ertl HA, Massey DJ, Hofley CM, Sohail F, Bienenstock EJ, Sebastian H, Chiolo I, LaRocque JR, Aging impairs double-strand break repair by homologous recombination in Drosophila germ cells, Aging Cell, 16 (2017) 320–328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [168].Pal S, Postnikoff SD, Chavez M, Tyler JK, Impaired cohesion and homologous recombination during replicative aging in budding yeast, Sci Adv, 4 (2018) eaaq0236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Richardson C, Stark JM, Ommundsen M, Jasin M, Rad51 overexpression promotes alternative double-strand break repair pathways and genome instability, Oncogene, 23 (2004) 546–553. [DOI] [PubMed] [Google Scholar]
  • [170].Wu Y, Xiao S, Zhu XD, MRE11-RAD50-NBS1 and ATM function as co-mediators of TRF1 in telomere length control, Nat Struct Mol Biol, 14 (2007) 832–840. [DOI] [PubMed] [Google Scholar]
  • [171].Tarsounas M, Munoz P, Claas A, Smiraldo PG, Pittman DL, Blasco MA, West SC, Telomere maintenance requires the RAD51D recombination/repair protein, Cell, 117 (2004) 337–347. [DOI] [PubMed] [Google Scholar]
  • [172].Yu EY, Kojic M, Holloman WK, Lue NF, Brh2 and Rad51 promote telomere maintenance in Ustilago maydis, a new model system of DNA repair proteins at telomeres, DNA Repair (Amst), 12 (2013) 472–479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [173].Badie S, Escandell JM, Bouwman P, Carlos AR, Thanasoula M, Gallardo MM, Suram A, Jaco I, Benitez J, Herbig U, Blasco MA, Jonkers J, Tarsounas M, BRCA2 acts as a RAD51 loader to facilitate telomere replication and capping, Nat Struct Mol Biol, 17 (2010) 1461–1469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [174].Mao P, Liu J, Zhang Z, Zhang H, Liu H, Gao S, Rong YS, Zhao Y, Homologous recombination-dependent repair of telomeric DSBs in proliferating human cells, Nat Commun, 7 (2016) 12154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [175].Jaco I, Munoz P, Goytisolo F, Wesoly J, Bailey S, Taccioli G, Blasco MA, Role of mammalian Rad54 in telomere length maintenance, Mol Cell Biol, 23 (2003) 5572–5580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [176].Feretzaki M, Pospisilova M, Valador Fernandes R, Lunardi T, Krejci L, Lingner J, RAD51-dependent recruitment of TERRA lncRNA to telomeres through R-loops, Nature, 587 (2020) 303–308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [177].Bettin N, Oss Pegorar C, Cusanelli E, The Emerging Roles of TERRA in Telomere Maintenance and Genome Stability, Cells, 8 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [178].Azzara A, Chiaramonte A, Filomeni E, Pinto B, Mazzoni S, Piaggi S, Angela Guzzardi M, Bruschi F, Iozzo P, Scarpato R, Increased level of DNA damage in some organs of obese Zucker rats by gamma-H2AX analysis, Environ Mol Mutagen, 58 (2017) 477–484. [DOI] [PubMed] [Google Scholar]
  • [179].Scarpato R, Verola C, Fabiani B, Bianchi V, Saggese G, Federico G, Nuclear damage in peripheral lymphocytes of obese and overweight Italian children as evaluated by the gamma-H2AX focus assay and micronucleus test, FASEB J, 25 (2011) 685–693. [DOI] [PubMed] [Google Scholar]
  • [180].Sidorova JM, Roles of the Werner syndrome RecQ helicase in DNA replication, DNA Repair (Amst), 7 (2008) 1776–1786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [181].Goto M, Miller RW, Ishikawa Y, Sugano H, Excess of rare cancers in Werner syndrome (adult progeria), Cancer Epidemiol Biomarkers Prev, 5 (1996) 239–246. [PubMed] [Google Scholar]
  • [182].Mukherjee S, Sinha D, Bhattacharya S, Srinivasan K, Abdisalaam S, Asaithamby A, Werner Syndrome Protein and DNA Replication, Int J Mol Sci, 19 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [183].Fang EF, Hou Y, Lautrup S, Jensen MB, Yang B, SenGupta T, Caponio D, Khezri R, Demarest TG, Aman Y, Figueroa D, Morevati M, Lee HJ, Kato H, Kassahun H, Lee JH, Filippelli D, Okur MN, Mangerich A, Croteau DL, Maezawa Y, Lyssiotis CA, Tao J, Yokote K, Rusten TE, Mattson MP, Jasper H, Nilsen H, Bohr VA, NAD(+) augmentation restores mitophagy and limits accelerated aging in Werner syndrome, Nat Commun, 10 (2019) 5284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [184].Verdin E, NAD(+) in aging, metabolism, and neurodegeneration, Science, 350 (2015) 1208–1213. [DOI] [PubMed] [Google Scholar]
  • [185].Rossi ML, Ghosh AK, Bohr VA, Roles of Werner syndrome protein in protection of genome integrity, DNA Repair (Amst), 9 (2010) 331–344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [186].Dhillon KK, Sidorova J, Saintigny Y, Poot M, Gollahon K, Rabinovitch PS, Monnat RJ Jr., Functional role of the Werner syndrome RecQ helicase in human fibroblasts, Aging Cell, 6 (2007) 53–61. [DOI] [PubMed] [Google Scholar]
  • [187].Bacolla A, Wang G, Jain A, Chuzhanova NA, Cer RZ, Collins JR, Cooper DN, Bohr VA, Vasquez KM, Non-B DNA-forming sequences and WRN deficiency independently increase the frequency of base substitution in human cells, J Biol Chem, 286 (2011) 10017–10026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [188].Damerla RR, Knickelbein KE, Strutt S, Liu FJ, Wang H, Opresko PL, Werner syndrome protein suppresses the formation of large deletions during the replication of human telomeric sequences, Cell Cycle, 11 (2012) 3036–3044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [189].Crabbe L, Verdun RE, Haggblom CI, Karlseder J, Defective telomere lagging strand synthesis in cells lacking WRN helicase activity, Science, 306 (2004) 1951–1953. [DOI] [PubMed] [Google Scholar]
  • [190].Sun L, Nakajima S, Teng Y, Chen H, Yang L, Chen X, Gao B, Levine AS, Lan L, WRN is recruited to damaged telomeres via its RQC domain and tankyrase1-mediated poly-ADP-ribosylation of TRF1, Nucleic Acids Res, 45 (2017) 3844–3859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [191].Liu FJ, Barchowsky A, Opresko PL, The Werner syndrome protein suppresses telomeric instability caused by chromium (VI) induced DNA replication stress, PLoS One, 5 (2010) e11152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [192].Crabbe L, Jauch A, Naeger CM, Holtgreve-Grez H, Karlseder J, Telomere dysfunction as a cause of genomic instability in Werner syndrome, Proc Natl Acad Sci U S A, 104 (2007) 2205–2210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [193].Shamanna RA, Lu H, de Freitas JK, Tian J, Croteau DL, Bohr VA, WRN regulates pathway choice between classical and alternative non-homologous end joining, Nat Commun, 7 (2016) 13785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [194].Moore G, Knoblaugh S, Gollahon K, Rabinovitch P, Ladiges W, Hyperinsulinemia and insulin resistance in Wrn null mice fed a diabetogenic diet, Mech Ageing Dev, 129 (2008) 201–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [195].Massip L, Garand C, Turaga RV, Deschenes F, Thorin E, Lebel M, Increased insulin, triglycerides, reactive oxygen species, and cardiac fibrosis in mice with a mutation in the helicase domain of the Werner syndrome gene homologue, Exp Gerontol, 41 (2006) 157–168. [DOI] [PubMed] [Google Scholar]
  • [196].Yamada K, Ikegami H, Yoneda H, Miki T, Ogihara T, All patients with Werner's syndrome are insulin resistant, but only those who also have impaired insulin secretion develop overt diabetes, Diabetes Care, 22 (1999) 2094–2095. [DOI] [PubMed] [Google Scholar]
  • [197].Cogger VC, Svistounov D, Warren A, Zykova S, Melvin RG, Solon-Biet SM, O'Reilly JN, McMahon AC, Ballard JW, De Cabo R, Le Couteur DG, Lebel M, Liver aging and pseudocapillarization in a Werner syndrome mouse model, J Gerontol A Biol Sci Med Sci, 69 (2014) 1076–1086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [198].Flanagan M, Cunniff C, Bloom Syndrome, GeneReviews [Internet], (2006. [updated 2019]). [Google Scholar]
  • [199].Cunniff C, Bassetti JA, Ellis NA, Bloom's Syndrome: Clinical Spectrum, Molecular Pathogenesis, and Cancer Predisposition, Mol Syndromol, 8 (2017) 4–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [200].de Renty C, Ellis NA, Bloom's syndrome: Why not premature aging?: A comparison of the BLM and WRN helicases, Ageing Res Rev, 33 (2017) 36–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [201].Szekely AM, Bleichert F, Numann A, Van Komen S, Manasanch E, Ben Nasr A, Canaan A, Weissman SM, Werner protein protects nonproliferating cells from oxidative DNA damage, Mol Cell Biol, 25 (2005) 10492–10506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [202].Lu H, Fang EF, Sykora P, Kulikowicz T, Zhang Y, Becker KG, Croteau DL, Bohr VA, Senescence induced by RECQL4 dysfunction contributes to Rothmund-Thomson syndrome features in mice, Cell Death Dis, 5 (2014) e1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [203].Opresko PL, von Kobbe C, Laine JP, Harrigan J, Hickson ID, Bohr VA, Telomere-binding protein TRF2 binds to and stimulates the Werner and Bloom syndrome helicases, J Biol Chem, 277 (2002) 41110–41119. [DOI] [PubMed] [Google Scholar]
  • [204].Yankiwski V, Marciniak RA, Guarente L, Neff NF, Nuclear structure in normal and Bloom syndrome cells, Proc Natl Acad Sci U S A, 97 (2000) 5214–5219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [205].Barefield C, Karlseder J, The BLM helicase contributes to telomere maintenance through processing of late-replicating intermediate structures, Nucleic Acids Res, 40 (2012) 7358–7367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [206].Chu WK, Hanada K, Kanaar R, Hickson ID, BLM has early and late functions in homologous recombination repair in mouse embryonic stem cells, Oncogene, 29 (2010) 4705–4714. [DOI] [PubMed] [Google Scholar]
  • [207].Xue C, Daley JM, Xue X, Steinfeld J, Kwon Y, Sung P, Greene EC, Single-molecule visualization of human BLM helicase as it acts upon double- and single-stranded DNA substrates, Nucleic Acids Res, 47 (2019) 11225–11237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [208].Bugreev DV, Mazina OM, Mazin AV, Bloom syndrome helicase stimulates RAD51 DNA strand exchange activity through a novel mechanism, J Biol Chem, 284 (2009) 26349–26359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [209].Wang H, Li S, Zhang H, Wang Y, Hao S, Wu X, BLM prevents instability of structure-forming DNA sequences at common fragile sites, PLoS Genet, 14 (2018) e1007816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [210].Davies SL, North PS, Hickson ID, Role for BLM in replication-fork restart and suppression of origin firing after replicative stress, Nat Struct Mol Biol, 14 (2007) 677–679. [DOI] [PubMed] [Google Scholar]
  • [211].Keller C, Keller KR, Shew SB, Plon SE, Growth deficiency and malnutrition in Bloom syndrome, J Pediatr, 134 (1999) 472–479. [DOI] [PubMed] [Google Scholar]
  • [212].Zheng L, Jia J, Finger LD, Guo Z, Zer C, Shen B, Functional regulation of FEN1 nuclease and its link to cancer, Nucleic Acids Res, 39 (2011) 781–794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [213].Xu X, Shi R, Zheng L, Guo Z, Wang L, Zhou M, Zhao Y, Tian B, Truong K, Chen Y, Shen B, Hua Y, Xu H, SUMO-1 modification of FEN1 facilitates its interaction with Rad9-Rad1-Hus1 to counteract DNA replication stress, J Mol Cell Biol, 10 (2018) 460–474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [214].Zhao J, Wang G, Del Mundo IM, McKinney JA, Lu X, Bacolla A, Boulware SB, Zhang C, Zhang H, Ren P, Freudenreich CH, Vasquez KM, Distinct Mechanisms of Nuclease-Directed DNA-Structure-Induced Genetic Instability in Cancer Genomes, Cell Rep, 22 (2018) 1200–1210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [215].Kucherlapati M, Yang K, Kuraguchi M, Zhao J, Lia M, Heyer J, Kane MF, Fan K, Russell R, Brown AM, Kneitz B, Edelmann W, Kolodner RD, Lipkin M, Kucherlapati R, Haploinsufficiency of Flap endonuclease (Fen1) leads to rapid tumor progression, Proc Natl Acad Sci U S A, 99 (2002) 9924–9929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [216].Larsen E, Kleppa L, Meza TJ, Meza-Zepeda LA, Rada C, Castellanos CG, Lien GF, Nesse GJ, Neuberger MS, Laerdahl JK, William Doughty R, Klungland A, Early-onset lymphoma and extensive embryonic apoptosis in two domain-specific Fen1 mice mutants, Cancer Res, 68 (2008) 4571–4579. [DOI] [PubMed] [Google Scholar]
  • [217].Saharia A, Guittat L, Crocker S, Lim A, Steffen M, Kulkarni S, Stewart SA, Flap endonuclease 1 contributes to telomere stability, Curr Biol, 18 (2008) 496–500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [218].Brosh RM Jr., von Kobbe C, Sommers JA, Karmakar P, Opresko PL, Piotrowski J, Dianova I, Dianov GL, Bohr VA, Werner syndrome protein interacts with human flap endonuclease 1 and stimulates its cleavage activity, EMBO J, 20 (2001) 5791–5801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [219].Sidorova JM, Li N, Folch A, Monnat RJ Jr., The RecQ helicase WRN is required for normal replication fork progression after DNA damage or replication fork arrest, Cell Cycle, 7 (2008) 796–807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [220].Sharma S, Otterlei M, Sommers JA, Driscoll HC, Dianov GL, Kao HI, Bambara RA, Brosh RM Jr., WRN helicase and FEN-1 form a complex upon replication arrest and together process branchmigrating DNA structures associated with the replication fork, Mol Biol Cell, 15 (2004) 734–750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [221].Saharia A, Teasley DC, Duxin JP, Dao B, Chiappinelli KB, Stewart SA, FEN1 ensures telomere stability by facilitating replication fork re-initiation, J Biol Chem, 285 (2010) 27057–27066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [222].Muftuoglu M, Wong HK, Imam SZ, Wilson DM 3rd, Bohr VA, Opresko PL, Telomere repeat binding factor 2 interacts with base excision repair proteins and stimulates DNA synthesis by DNA polymerase beta, Cancer Res, 66 (2006) 113–124. [DOI] [PubMed] [Google Scholar]
  • [223].Sun L, Tan R, Xu J, LaFace J, Gao Y, Xiao Y, Attar M, Neumann C, Li GM, Su B, Liu Y, Nakajima S, Levine AS, Lan L, Targeted DNA damage at individual telomeres disrupts their integrity and triggers cell death, Nucleic Acids Res, 43 (2015) 6334–6347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [224].Tyynismaa H, Sembongi H, Bokori-Brown M, Granycome C, Ashley N, Poulton J, Jalanko A, Spelbrink JN, Holt IJ, Suomalainen A, Twinkle helicase is essential for mtDNA maintenance and regulates mtDNA copy number, Hum Mol Genet, 13 (2004) 3219–3227. [DOI] [PubMed] [Google Scholar]
  • [225].Spelbrink JN, Li FY, Tiranti V, Nikali K, Yuan QP, Tariq M, Wanrooij S, Garrido N, Comi G, Morandi L, Santoro L, Toscano A, Fabrizi GM, Somer H, Croxen R, Beeson D, Poulton J, Suomalainen A, Jacobs HT, Zeviani M, Larsson C, Human mitochondrial DNA deletions associated with mutations in the gene encoding Twinkle, a phage T7 gene 4-like protein localized in mitochondria, Nat Genet, 28 (2001) 223–231. [DOI] [PubMed] [Google Scholar]
  • [226].Peter B, Falkenberg M, TWINKLE and Other Human Mitochondrial DNA Helicases: Structure, Function and Disease, Genes (Basel), 11 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [227].Tyynismaa H, Mjosund KP, Wanrooij S, Lappalainen I, Ylikallio E, Jalanko A, Spelbrink JN, Paetau A, Suomalainen A, Mutant mitochondrial helicase Twinkle causes multiple mtDNA deletions and a late-onset mitochondrial disease in mice, Proc Natl Acad Sci U S A, 102 (2005) 17687–17692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [228].Remtulla S, Emilie Nguyen CT, Prasad C, Campbell C, Twinkle-Associated Mitochondrial DNA Depletion, Pediatr Neurol, 90 (2019) 61–65. [DOI] [PubMed] [Google Scholar]
  • [229].Sarzi E, Goffart S, Serre V, Chretien D, Slama A, Munnich A, Spelbrink JN, Rotig A, Twinkle helicase (PEO1) gene mutation causes mitochondrial DNA depletion, Ann Neurol, 62 (2007) 579–587. [DOI] [PubMed] [Google Scholar]
  • [230].Korhonen JA, Pham XH, Pellegrini M, Falkenberg M, Reconstitution of a minimal mtDNA replisome in vitro, EMBO J, 23 (2004) 2423–2429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [231].Foote K, Reinhold J, Yu EPK, Figg NL, Finigan A, Murphy MP, Bennett MR, Restoring mitochondrial DNA copy number preserves mitochondrial function and delays vascular aging in mice, Aging Cell, 17 (2018) e12773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [232].Mengel-From J, Thinggaard M, Dalgard C, Kyvik KO, Christensen K, Christiansen L, Mitochondrial DNA copy number in peripheral blood cells declines with age and is associated with general health among elderly, Hum Genet, 133 (2014) 1149–1159. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES