Oxidatively Induced DNA Damage and Its Repair in

Miral Dizdaroglu

Biomolecular Measurement Division, National Institute of Standards and Technology, 100

Bureau Drive, MS 8311, Gaithersburg, MD 20899, USA

Corresponding author. Tel.: +1-301-975-2581; fax: +1-301-975-8505

E-mail address: [email protected]

.

1 ABSTRACT

Oxidatively induced DNA damage is caused in living organisms by endogenous and exogenous reactive species. DNA lesions resulting from this type of damage are mutagenic and cytotoxic and, if not repaired, can cause genetic instability that may lead to disease processes including carcinogenesis. Living organisms possess DNA repair mechanisms that include a variety of pathways to repair multiple DNA lesions.

Mutations and polymorphisms also occur in DNA repair genes adversely affecting

DNA repair systems. Cancer tissues overexpress DNA repair proteins and thus develop

greater DNA repair capacity than normal tissues. Increased DNA repair in tumors that

removes DNA lesions before they become toxic is a major mechanism for development

of resistance to therapy, affecting patient survival. Accumulated evidence suggests that

DNA repair capacity may be a predictive biomarker for patient response to therapy.

Thus, knowledge of DNA protein expressions in normal and cancerous tissues may

help predict and guide development of treatments and yield the best therapeutic

response. DNA repair proteins constitute targets for inhibitors to overcome the

resistance of tumors to therapy. Inhibitors of DNA repair for combination therapy or as

single agents for monotherapy may help selectively kill tumors, potentially leading to

personalized therapy. Numerous inhibitors have been developed and are being tested in

clinical trials. The efficacy of some inhibitors in therapy has been demonstrated in

patients. More developments of inhibitors of DNA repair proteins are globally

underway to help eradicate cancer.

2

Keywords:

Cancer therapy

DNA damage

DNA repair

DNA glycosylases

Inhibitors

3

Contents

1. Introduction

2. Mechanistic aspects of oxidatively induced DNA damage

2.1. Purines

2.2. Pyrimidines

2.3. Sugar moiety

2.4. Tandem lesions

2.4.1. 8,5'-Cyclopurine-2'-deoxynucleosides

2.4.2. Base-base tandem lesions

2.4.3. DNA-protein cross-links

2.4.4. Clustered lesions

3. Cellular repair of oxidatively induced DNA lesions

3.1.

3.1.1. Substrate specificities of prokaryotic DNA glycosylases

3.1.2. Substrate specificities of eukaryotic DNA glycosylases

3.1.3. Repair of sugar lesions

4. Genetic effects of oxidatively induced DNA lesions

4.1. Purine-derived lesions

4.2. Pyrimidine-derived lesions

4.3. 8,5'-Cyclopurine-2'-deoxynucleosides

4.4. Sugar lesions

5. Oxidatively induced DNA damage and cancer

5.1. Role of DNA glycosylases of BER in carcinogenesis

5.1.1. OGG1

5.1.2. NEIL proteins

4

5.1.3. NTH1

5.2. Other BER proteins

5.2.1. APE1

5.2.2. Pol β

5.3. DNA lesions and DNA repair proteins as biomarkers

5.3.1. DNA lesions as biomarkers

5.3.2. BER proteins as biomarkers

5.3.3. BER proteins as therapy targets

6. Conclusions

References

5

● ●─ ─ Abbreviations: RS, reactive species; OH, hydroxyl radical; O2 , superoxide radical; eaq ,

hydrated electron; k, reaction rate constant; 8-OH-Gua, 8-hydroxyguanine; FapyGua, 2,6- diamino-4-hydroxy-5-formamidopyrimidine; 8-OH-Ade, 8-hydroxyadenine; FapyAde, 4,6- diamino-5-formamidopyrimidine; Sp, spiroiminohydantoin; Gh, 5-guanidinohydantoin; 5-

OHMe-Ura, 5-(hydroxymethyl)uracil; 5-OH-Cyt, 5-hydroxycytosine; 5-OH-Ura, 5- hydroxyuracil; cdA, 8,5'-cyclo-2'-deoxyadenosine; cdG, 8,5'-cyclo-2'-deoxyguanosine; Fo, formamido residue; Thy-Tyr cross-link, 3-[1,3-dihydro-2,4-dioxopyrimidin-5-yl)-methyl]-L- tyrosine; BER, base excision repair; NER, excision repair; MMR, mismatch repair; AP site, apyrimidinic/apurinic site; Me-FapyGua, 2,6-diamino-4-hydroxy-N7-methyl-

5-formamidopyrimidine; APE1, apurinic/apyrimidinic endonuclease 1; dRP, 2'-deoxyribose phosphate; Pol β; DNA polymerase β; TS, thymidylate synthetase.

6

1. Introduction

Reactive species (RS) including free radicals derived from either oxygen or nitrogen are generated in aerobic organisms by cellular metabolism and by exogenous sources such as ionizing radiations, UV radiation, redox cycling drugs, carcinogenic compounds, environmental toxins, etc. [1]. Antioxidant defense mechanisms exist in living organisms to encounter the production and effects of RS. If the prooxidant-antioxidant balance is disturbed in favor of the former, a state of oxidative stress can occur, leading to oxidative damage to biomolecules including DNA, proteins and lipids [1,2]. Consequences of oxidative stress can

be manyfold depending on its severity and the cell type [1]. Among others, these may include

increased genetic instability, proliferation, reduction of antioxidants, , and

angiogenesis [1]. Oxidative stress can also drive the onset of inflammation, which produces

RS and is a hallmark of cancer, predisposing individuals to different types of [3-5].

The acute inflammatory response recruits neutrophils that can damage DNA [4,6]. Reactive

species are involved in carcinogenesis by damaging DNA and by modulating certain cellular

pathways[1,4]. These species can be radicals or non-radicals. Among the oxygen-derived

radicals, the hydroxyl radical (●OH) is the most reactive one and reacts with biological

molecules such as DNA constituents at or near diffusion-controlled rates [7]. Other radicals

●─ ● ● such as superoxide radical (O2 ), hydroperoxyl radical (HO2 ), peroxyl radicals (RO2 ),

● 1 + alkoxyl radicals (RO ) and singlet oxygen (O2 Σg ) possess very low or intermediate

reactivity. Non-radical H2O2 is not reactive, unless its reaction with transition metal ions

converts it into ●OH [1]. Nitric oxide (NO●) is also a free radical and possesses low

●─ reactivity; however, its reaction with O2 is diffusion-controlled, yielding peroxynitrite

(ONOO─) [8]. Peroxynitrite is a fairly unreactive non-radical. On the other hand, its

protonated form peroxynitrous acid (ONOOH) can undergo homolytic fission to yield ●OH

● and NO2 , although this reaction may not be favored [1]. Ionizing radiations also generate

7

● ● ─ OH and, in addition, H atom (H ) (also a free radical) and hydrated electron (eaq ) from

cellular water [9]. Reactions of these endogenously and exogenously generated species with

the DNA bases and sugar moiety result in the formation of a multitude of modifications

(reviewed in [9,10]). This type of damage, which is called oxidatively induced DNA damage, can be repaired in living cells by a variety of repair mechanisms [11]. Oxidatively induced

DNA modifications that escape repair before replication may lead to , which is

well known to be a fundamental part of the molecular basis of all cancers [11-13].

occur throughout the genome, including in genes that maintain genetic stability, leading to

genetic instability, which is a hallmark of cancer [3,14,15]. Genetic instability may affect

many types of enzymes in various pathways including DNA repair [11]. In healthy cells,

there is a balance between DNA damage and DNA repair. In cancer cells, however, this

balance may be disturbed in favor of DNA damage, overwhelming DNA repair capacity of

cells and thus resulting in mutations at high frequency. Increase in DNA repair capacity may

also occur in cancer cells, causing therapy resistance [16-19]. There is mounting evidence

that oxidatively induced DNA damage by endogenous and exogenous sources may be a

significant source of mutations and genomic instability, and thus an important contributor to

carcinogenesis [11,20-22].

2. Mechanistic aspects of oxidatively induced DNA damage

2.1. Purines

Mechanistic aspects of oxidatively induced DNA damage has extensively been reviewed

in the past and just recently [9,10,23]. Thus, only a brief summary of this field will be given

here. Of the RS, ●OH is the most damaging species to DNA and other biological molecules.

Its reactions by addition to the double bonds of purines and pyrimidines in DNA are

diffusion-controlled with second-order reaction rate constants (k) of 4‒9 x 109 dm3 mol‒1 s‒1

8

[7,9]. Abstraction of H● from the five C-atoms of the sugar moiety and from the methyl group of also occurs, albeit by slower rates with k ≈ 2 x 109 M‒1 s‒1 [7,9]. Ionizing

─ radiation-generated eaq also adds to the double bonds of DNA bases at diffusion-controlled

rates (k = 0.9‒1.7 x 109 dm3 mol‒1 s‒1); however, the addition reactions of H● are slower, but

8 3 ‒1 ‒1 ─ ● still occur at appreciable rates (k = 1‒5 x 10 dm mol s ) [23-27]. Reactions of eaq and H

with the sugar moiety of DNA are negligible. Hydroxyl radical preferentially adds to the sites

of double bonds of purines and pyrimidines with the highest electron density because of its

electrophilic nature. Addition to guanine generates C4-OH‒, C5-OH‒ and C8-OH‒adduct

● ● radicals [10,23,28]. In addition, an H abstraction by OH from the NH2 group at C2 of

guanine has been claimed to occur at a rate of ≈ 65%, practically eliminating the addition of

●OH to C4 [29,30]. However, theoretical and experimental studies unequivocally showed that this reaction is energetically not favored and does not occur to an appreciable extent [9,10,31-

36]. Addition reactions of ●OH with adenine mainly produces C4-OH‒ and C8-OH‒adduct radicals, although the C5-OH‒adduct radical is also formed, but to a much lesser extent

[23,37,38]. The addition of ●OH to the C2 of adenine also occurs to an extent of 2%. The C4-

OH‒ and C5-OH‒adduct radicals of guanine and adenine dehydrate to yield Gua(‒H)● and

Ade(‒H)● radicals, respectively, that may be reduced to reconstitute Gua and Ade [23,39].

Gua(‒H)● protonates to give rise to guanine radical cation (Gua●+), which can be converted

into the C8-OH‒adduct radical upon hydration (HO‒ addition) [40-42]. Direct effect of

ionizing radiation also generates Gua●+; therefore, the direct effect and the indirect effect of

ionizing radiation may lead to the same products of guanine [43,44]. Ade(‒H)● may also

protonate to give Ade●+, which may yield the C8-OH‒adduct radical upon hydration. This is

the same adduct radical formed by direct addition of ●OH to the C8 of adenine. The addition of ●OH to the C2 of adenine also occurs to an extent of 2%.

9

The major products of guanine in DNA result from the reactions of the C8-OH‒adduct radical, one-electron oxidation of which gives rise to 8-hydroxyguanine (8-OH-Gua) [45-50].

In an exothermic reaction, 8-OH-Gua tautomerizes into 8-oxoguanine, which is its

predominant keto form [51-53]. In contrast to oxidation, the C8-OH‒adduct radical of

guanine can undergo a β-fragmentation leading to unimolecular opening of the imidazole ring

(k = 2 x 105 s‒1) [23,39,54,55], followed by one-electron reduction to yield 2,6-diamino-4- hydroxy-5-formamidopyrimidine (FapyGua) (reviewed in [10,23]). The formation of 8-OH-

Gua increases in the presence of O2, whereas ring-opening leading to FapyGua is favored at low O2 concentrations and can compete with the bimolecular oxidation. Under hypoxic conditions of the cell nucleus, therefore, the ring-opening of the C8-OH‒adduct radical followed by reduction may be a favorable reaction. This notion is strongly supported by the formation of FapyGua in living cells with comparable yields to that of 8-OH-Gua (reviewed in [56]). Adenine undergoes analogous reactions, generating 8-hydroxyadenine (8-OH-Ade) and 4,6-diamino-5-formamidopyrimidine (FapyAde) (reviewed in [56]. Although called

pyrimidines, FapyGua and FapyAde are distinguished from pyrimidines by the position of

their glycosidic bond attached to the sugar moiety in DNA through the amino group at C6 of

the pyrimidine ring. It should be emphasized that, chemically and mechanistically, and also in

terms of biological effects, these formamidopyrimidines are different from their methylated

counterparts (reviewed in [56]). The oxidation of the C2-OH‒adduct radical of adenine

results in the formation of 2-hydroxyadenine [57,58].

Oxygen reacts with the OH‒adduct radicals of guanine and adenine at different rates. The

6 3 C4-OH‒adduct radical of guanine does not react with O2 at appreciable rates (k ≤ 10 dm

‒1 ‒1 9 mol s ); however, O2 rapidly reacts with the C4-OH‒adduct radical of adenine (k = 1 x 10

3 ‒1 ‒1 dm mol s )[38]. On the other hand, the reactions of O2 with the C8-OH‒adduct radicals of

guanine and adenine are diffusion-controlled (k = 4 x 109 dm3 mol‒1 s‒1) [38,39]. This is

10 likely to be the reason for the preferred formation of 8-OH-Gua and 8-OH-Ade in the

presence of oxygen. The ring-opening leading to formamidopyrimidines, however, may

compete with this reaction and thus be equally efficient at the hypoxic conditions of the cell

nucleus. The abundant formation of formamidopyrimidines in vivo supports this notion

● (reviewed in [10]). The reaction of O2 with Gua(‒H) has been reported to lead to the

formation of 2,5-diamino-4H-imidazol-4-one and 2,2,4-triamino-5(2H)-oxazolone [59,60].

●─ ● However, this mechanism has not been confirmed. Instead, the addition of O2 to Gua(–H) has been shown to be kinetically more favored reaction in nucleosides and DNA (k = 3–4.7 x

109 dm3 mol‒1 s‒1) [39,61,62]. This reaction generates guanine hydroperoxide that undergoes several reactions to give rise to 2,5-diamino-4H-imidazol-4-one, the slow hydrolysis of which results in the formation of 2,2,4-triamino-5(2H)-oxazolone [59,62-64].

Because of its low reduction potential (0.74 V) compared to that of guanine (1.29 V), 8-

OH-Gua is even more prone to oxidation than guanine by a number of oxidizing agents

2─ including ionizing radiation, metal ions, peroxynitrate and IrCl6 [65]. Its oxidation gives

rise to 8-OH-Gua●+, which readily hydrates (HO─ addition) and produces the 5-OH–adduct

radical of 8-OH-Gua. One-electron oxidation of this radical leads to 5-OH-8-OH-Gua, the

isomerization (acyl shift) of which gives rise to spiroiminohydantoin (Sp) and 5-

guanidinohydantoin (Gh) by loss of CO2 depending on reaction conditions [66-72]. Cadet et

al. have misassigned the structure of spiroiminohydantoin as 4,8-dihydro-4-hydroxy-8- oxoguanine for almost two decades and have used it for a marker of single oxygen-induced damage to Gua [73-77]. Later on, the correct structure of this compound as spiroiminohydantoin has been elucidated using the synthesized authentic material and a number of analytical techniques [66-70]. Single oxygen also reacts with 8-OH-Gua to give rise to oxaluric acid and parabanic acid among other products [78,79]. Moreover, the reaction

●─ ●+ ● 9 3 ‒1 of O2 with 8-OH-Gua and its deprotonated form 8-OH-Gua(‒H ) (k = 3 x 10 dm mol

11 s‒1) yields 5-hydroperoxide of 8-OH-Gua whose facile decomposition leads to the formation

of oxaluric acid and parabanic acid [80]. This area has extensively been reviewed elsewhere

[72,81]. All these data unequivocally show the possible effect of numerous factors on 8-OH-

Gua leading to its decomposition, and thus on its measured level in DNA. It should be

pointed out that this mounting evidence stands in stark contrast to the claim by the European

Standards Committee on Oxidative DNA Damage (ESCODD) about the so-called “correct” or

“established” value of the background level of 8-OH-Gua in living organisms and about the

validity of its related recommendations to editors and reviewers of manuscripts not to accept

manuscript with levels of 8-OH-Gua above the “established” value [82-84].

─ Ionizing radiation-generated eaq reacts with guanine and adenine at diffusion-controlled

rates, giving rise to radical anions Gua●─ and Ade●─, respectively (k = 3.3‒6 x 109 dm3 mol‒1

s‒1) [25-27]. These radical anions rapidly protonate in reaction with water, generating Gua(‒

H●) and Ade(‒H●), which subsequently yield the C8-H‒adduct radicals of guanine and

adenine, respectively, by water-assisted tautomerization [23,26,27,85-87]. Addition of H● at

C8 of guanine and adenine also generates the C8-H‒adduct radicals [26]. No products of

these adduct radicals have been identified in DNA. This is likely due to electron transfer from

these radicals to other DNA bases prior to formation of final products [87].

2.2. Pyrimidines

Hydroxyl radical reacts with thymine and cytosine at diffusion control rates (k = 6.4‒6.8 x

109 dm3 mol‒1 s‒1) by addition to the C5‒C6 double bonds producing C5-OH‒ and C6-OH‒

adduct radicals [25,88]. Abstraction of H● from the methyl group of thymine also occurs,

albeit to a lesser extent, giving rise to an allyl radical [89,90]. Addition of ●OH occurs more

at C5 than that at C6 because of the higher electron density at C5. Thymine and cytosine

radicals are oxidized or reduced depending on their redox properties, experimental conditions

12 and the presence or absence of oxygen, producing a variety of products (reviewed in [9,10]).

In the absence of oxygen, the C5-OH‒ and C6-OH‒adduct radicals of thymine and cytosine

undergo oxidation and reaction with water (HO─ addition) to yield thymine glycol (Thy

glycol) and cytosine glycol, respectively. 5-(Hydroxymethyl)uracil (5-OHMe-Ura) is formed by oxidation of the allyl radical of thymine followed by HO─ addition. The adduct radicals

are also reduced, producing 5-hydroxy-6-hydrothymine, 6-hydroxy-5-hydrothymine and 5-

hydroxy-6-hydrocytosine. Oxygen adds to the thymine and cytosine radicals at diffusion-

controlled rates generating peroxyl radicals (k ≈ 2 x 109 dm3 mol‒1 s‒1). 5-Hydroxy-6- hydrothymine, 6-hydroxy-5-hydrothymine and 5-hydroxy-6-hydrocytosine are thus not formed under oxygenated conditions. Further reactions of peroxyl radicals yield thymine glycol, 5-hydroxy-5-methylhydantoin, 5-OHMe-Ura, 5-formyluracil, cytosine glycol, dialuric acid, alloxan, 5-hydroxyhydantoin, isodialuric acid and 5,6-dihydroxycytosine. Cytosine products undergo dehydration and deamination to yield 5-hydroxycytosine (5-OH-Cyt), uracil glycol, 5-hydroxyuracil (5-OH-Ura) and 5-hydroxy-6-hydrouracil [91]. Hydrated electron reacts with thymine and cytosine by addition to 5,6-double bonds (k = 1.3‒1.8 x 1010

dm3 mol‒1 s‒1) giving rise to electron adducts (anion radicals). Addition of H● also takes

place, albeit at lower rates (k ≈ 1‒1.8 x 108 dm3 mol‒1 s‒1) and generates H-adduct radicals.

Protonation of anion radicals also produce H-adduct radicals, which are converted to 5,6- dihydrothymine and 5,6-dihydrocytosine by reduction. The deamination of the latter yields

5,6-dihydrouracil. In the presence of oxygen, these products are not formed because of

─ ● diffusion-controlled reactions of O2 with eaq and H , preventing the addition reactions of

these two species. However, 5,6-dihydropyrimidines may be formed in DNA in vivo because

of the hypoxic conditions of the cell nucleus.

Fig. 1 illustrates the main products of DNA bases. Many of these products were identified

in DNA in vitro, in cultured mammalian cells, and in animal and human tissues (reviewed in

13

[9,10,22,92]). Their types and yields depend on experimental conditions, the presence or absence of O2, cellular redox environment, disease conditions, DNA repair capacity,

scavenger concentration, among others.

2.3. Sugar moiety

Hydroxyl radical abstracts H● from all five carbons of 2'-deoxyribose in DNA (k = 2.5 x

109 dm3 mol‒1 s‒1) in the order of H5' > H4' > H3' ≈ H2' ≈ H1', leading to C-centered radicals

[9,25,93-95]. The order of abstraction follows the exposure to solvent with the C4'- and the

C5'-positions being the most accessible to solvent and from the minor groove. Reactions of

─ ● ● eaq and H are negligible. The extent of OH attack on 2'-deoxyribose in DNA may amount to ≈ 20%, although this ratio may depend on the cellular environment[9]. Further reactions of

the C-centered radicals of 2'-deoxyribose in the presence or absence of O2 cause DNA strand

breaks and release of intact DNA bases, and generate products that are either freed from

DNA or remained within DNA or are bound to DNA as end groups of broken DNA strands.

In 1970s, the oxidatively induced products of 2'-deoxyribose in DNA have been identified and reaction mechanisms of product formation and DNA strand breaks have been elucidated

[96-101]. The reactions resulting from the C4' radical in the absence of O2, leading to strand

breaks and products, were the first understood mechanistically [9,96]. These reactions still

remain the most-widely studied mechanism of product formation and strand breakage in

DNA. Fig. 2 illustrates the products of 2'-deoxyribose in DNA. Extensive reviews of the

mechanisms and product formation can be found elsewhere [9,10,102,103].

2.4. Tandem lesions

2.4.1. 8,5'-Cyclopurine-2'-deoxynucleosides

The H● abstraction by ●OH from C5' of 2'-deoxyribose causes the formation of the tandem

lesions 8,5'-cyclopurine-2'-deoxynucleosides in DNA. The stereospecific attack of the C5'-

14 centered radical at the C8 of purine nucleosides leads to C5'‒C8-intramolecular cyclization and an N-centered purine radical. The oxidation of this radical causes the formation of (5'R)-

and (5'S)-8,5'-cyclopurine-2'-deoxynucleosides. This reaction has been first discovered to

take place in adenosine-5'-monophosphate [104]. Subsequent studies showed that this

reaction also occurs in DNA generating both (5'R)- and (5'S)-diastereomers of 8,5'-cyclo-2'-

deoxyadenosine (R-cdA and S-cdA) and 8,5'-cyclo-2'-deoxyguanosine (R-cdG and S-cdG)

(reviewed in [105]). Fig. 1 illustrates the structures of these compounds. The C5'‒C8-

intramolecular cyclization is inhibited by O2 because of its rapid reaction with C-centered

9 3 ‒1 ‒1 radicals (k = 1.9 x 10 dm mol s ) [9,105,106]. At low O2 concentrations, however, 8,5'-

cyclopurine-2'-deoxynucleosides are formed, suggesting that a competition takes place

between the C5'‒C8-intramolecular cyclization and the reaction of O2 with the C5'-radical

[105,107]. This competition may also occur in living cells because of hypoxic conditions in the cell nucleus and steric hindrances. Indeed, R-cdA, S-cdA, R-cdG and S-cdG have been observed in cultured mammalian cells, and in human and animal tissues at background levels or at elevated levels depending on disease states, aging, DNA repair deficiency, gene knock- outs, environmental pollutants or exposure to ionizing radiation [108-126]. R-cdA and S-cdA have also been detected in human urine [127]. Moreover, these compounds have been found in urine of atherosclerosis patients at significantly greater concentrations than in that of healthy individuals [128]. These findings suggested that R-cdA and S-cdA combined with the noninvasive nature of urine collection may be used as potential disease biomarkers for basic research, and for clinical and epidemiological studies. The identification of the 8,5'- cyclopurine-2'-deoxynucleosides in human and animal tissues at background levels or after exposure to DNA-damaging agents, and in human urine is in stark contrast to a claim that the

“estimated” levels of these compounds in vivo were too low; therefore, they would not be detectable by any means, unless very high and biologically irrelevant ionizing radiation doses

15 were used [92,129-131]. It should be emphasized again that the background levels of R-cdA and S-cdA (or R-cdG and S-cdG) in cells could not be measured and, interestingly, were only estimated, without providing any data to support this claim. Moreover, human monocytes in culture and biologically irrelevant high radiation doses have been used only, and no data on human or animal tissues have been provided unlike the studies cited above. These facts make the aforementioned claim baseless and unscientific.

2.4.2. Base-base tandem lesions

Adjacent, intrastrand and interstrand base-base tandem lesions have been identified mostly in vitro in oligodeoxynucleotides and DNA upon exposure to ionizing radiation or to other

●OH-generating agents. It is out of the scope of this paper to review all the work done in this field and cite all the references. Briefly, identified adjacent tandem lesions in

oligodeoxynucleotides and DNA consisted of an 8-OH-Gua residue and a formamido residue

(Fo) as 8-OH-Gua/Fo or Fo/8-OH-Gua [132-143]. A mechanism has been proposed, which

involves the one-electron oxidation of a neighboring Gua by the C5-OH-C6-peroxyl radical of Thy followed by hydration of Gua●+ and oxidation to form 8-OH-Gua and the decomposition of the C5-OH-C6-oxyl radical of Thy yielding Fo [141]. However, the

significantly lower reduction potential of a peroxyl radical than that of Gua renders this

reaction endothermic. Thus, this mechanism has been dismissed as a very unlikely one [9].

These tandem lesions (8-OH-Gua/Fo or Fo/8-OH-Gua) have not yet been identified in

cellular DNA [92]. Intrastrand cross-links between C8 of guanine or adenine and the methyl

group of thymine have been observed as Gua[8,5-Me]Thy and Thy[5-Me,8]Gua [138,144-

149], or Ade[8,5-Me]Thy and Thy[5-Me,8]Ade [145,150]. Intrastrand cross-linking between guanine and cytosine (Gua[8,5]Cyt), guanine and 5-methylcytosine (Gua[8,5-Me]MeCyt), and guanine and thymine (Gua[8,N3]Thy) have also been reported [151-154]. An interstrand cross-link has been shown to occur between the amino group of adenine on one DNA strand

16 and the allyl radical of thymine on the other DNA strand [155-158]. In studies in vivo, however, only Gua[8,5-Me]Thy and Gua[8,5]Cyt have been identified in γ-irradiated cultured cells and in animal tissues [151,159-161]. Fig. 3 illustrates the structures of Gua[8,5-Me]Thy and Gua[8,5]Cyt. Extensive reviews including reaction mechanisms of this field can be found elsewhere [10,92,149].

2.4.3. DNA-protein cross-links

Hydroxyl radical reactions with DNA bases and proteins in chromatin, which generate

DNA base radicals and amino acid radicals, cause the formation of covalent DNA-protein cross-linking [162-168]. Mechanisms of cross-linking may involve the reaction between a

DNA base radical and an amino acid or an amino acid radical and a DNA base or a DNA base radical and an amino acid radical. For example, a Thy-Tyr cross-link (3-[1,3-dihydro-

2,4-dioxopyrimidin-5-yl)-methyl]-L-tyrosine) has been identified in γ-irradiated mixtures of

Thy with Tyr or with a peptide containing Tyr, and its structure has been elucidated [169-

174]. The structure of this DNA-protein cross-link is illustrated in Fig. 3. Subsequently, Thy-

Tyr cross-links have been detected in mammalian chromatin upon exposure to γ-irradiation or by treatment with H2O2/iron or copper ions [175,176]. Its formation has been proposed to result from the addition of the allyl radical of Thy (see above) to C3 of Tyr, followed by oxidation or from the combination of the allyl radical of Thy with the phenoxyl radical of

Tyr. The latter is formed by ●OH addition to the Tyr ring followed by water elimination

[177]. The cross-linking was not inhibited by O2, most likely because the allyl radical of Thy

adds to Tyr in close proximity without first reacting with O2. The Thy-Tyr cross-link has also been observed in mammalian cells upon exposure to ionizing radiation, H2O2 or Fe(II)-ions

[178,179], and in renal chromatin of rats upon treatment with a renal carcinogen [180].

Additional DNA-protein cross-links have been identified in γ-irradiated mammalian chromatin in vitro between Thy and Gly, Ala, Val, Leu, Ile, Thr and Lys, and between Cyt

17 and Tyr [175,181,182]. A lysine-guanine cross-linking involving a guanine radical cation has been observed in aerated aqueous solution of a thymine-guanine-thymine oligodeoxynucleotide in the presence of a trilysine peptide due to riboflavin-mediated photosensitization; however, the corresponding Nϵ-(guanine-8-yl)-lysine cross-link has not been identified in mammalian chromatin [183].

2.4.4. Clustered lesions

Clustered lesions in DNA are also known as locally multiply damaged sites and are produced almost exclusively by ionizing radiations [184-189]. These lesions can be formed on the same strand or on opposite strands within one or two helical turns of DNA and can persist in living cells due to resistance to DNA repair by DNA glycosylases or by endonucleases. This field has extensively been reviewed [190].

3. Cellular repair of oxidatively induced DNA lesions

Living organisms evolved to possess DNA repair mechanisms to repair DNA damage and thus to protect the genetic stability for survival (reviewed in [11,191,192]). Failure to repair

DNA damage may lead to detrimental biological consequences for organisms. There are numerous DNA repair mechanisms. Oxidatively induced DNA lesions are generally repaired by base excision repair (BER) and, to a lesser extent, by nucleotide excision repair (NER) both of which include multiple steps and enzymes (reviewed in [11]). DNA lesions paired with a cognate DNA base are repaired by mismatch repair (MMR) [193-196]. There is repair in the nucleotide pool where modified 2'-deoxynucleoside triphosphates are dephosporylated by MutT in E.coli and its homolog MTH1 in human and other mammalian cells, and are thus prevented from being incorporated into DNA by DNA polymerases during DNA replication

[197-202]. MTH1 has been found to be overexpressed in many cancers [203]. Moreover, cancer cells have been shown to require MTH1 for efficient survival, suggesting that this

18 protein may be targeted as an anticancer therapeutic approach [204]. DNA single-strand breaks are repaired by mechanisms similar to those in BER, whereas or non-homologous end-joining mechanisms act on double-strand breaks

(reviewed in [11,205,206]).

3.1. Base excision repair

Base excision repair is highly conserved during evolution from bacteria to humans. It starts

with the removal (excision) of a DNA base lesion from DNA by a DNA glycosylase that

hydrolyzes the N-glycosidic bond between the sugar moiety and the modified base, leaving

behind an abasic site, also called an apyrimidinic/apurinic (AP) site. DNA glycosylases are

either monofunctional removing the DNA lesion only or possess an associated AP-lyase

activity. The 3'-phosphodiester bond of the AP site is hydrolyzed by a β- or β-δ-elimination mechanism that generates a strand break with a 3' α,β-unsaturated (β-elimination) or a 5'-phosphate group (β-δ-elimination) [11]. Subsequently, AP-endonucleases, DNA polymerases and DNA ligases process AP sites to restore the DNA structure. The lyase activity is generally associated with DNA glycosylases specific for oxidatively induced DNA base lesions. BER consists of short-patch and long-patch pathways. The former is initiated by a bifunctional DNA glycosylase, whereas a monofunctional DNA glycosylase can start either pathway [207]. In general, the short-patch pathway is initiated by DNA glycosylases, whereas

AP sites resulting from spontaneous hydrolysis of DNA bases are repaired by the long-patch pathway [208,209]. DNA glycosylases are highly conserved from bacteria to humans and are divided into two families, the Nth superfamily and the Fpg/Nei family, on the basis of structure and sequence homology [208,210-212]. The members of the Nth superfamily are widely found in bacteria, archaea and , whereas those of the Fpg/Nei family are sparsely distributed across the phyla. All members of these families use a common

19 mechanism for catalysis that includes several steps [213,214]. DNA glycosylases in the

Fpg/Nei family are characterized by a helix-two turn-helix (H2TH) motif and a zinc or zincless finger motif for DNA binding. They also have a conserved N-terminus with a Pro residue, which is essential for catalysis. The Fpg/Nei family members are named after bacterial members formamidopyrimidine glycosylase (Fpg, also called MutM) and endonuclease VIII (Nei), and also include NEIL1, NEIL2 and NEIL3; the Nth superfamily contains E. coli endonuclease III (Nth), E. coli MutY, yeast Ntg1 and Ntg2, MUTYH, OGG1 and AlkA [215].

DNA glycosylases have distinct substrate specificities, although there exists a redundancy with respect to overlapping substrates. The determination of substrate specificities of DNA glycosylases has been performed using various substrates, methods and techniques. In general, most studies used oligodeoxynucleotides with a single DNA lesion incorporated at a specific position. The use of such substrates and applied analytical methods permitted to study the excision of a single modified DNA base only at a time. A different concept that uses damaged DNA with multiple lesions and the technique of gas chromatography-mass spectrometry (GC-MS) has been proposed for the determination of substrate specificities and excision kinetics of DNA glycosylases [49]. This technique permits the simultaneous

identification and quantification of multiple modified bases from all four DNA bases in a

given DNA sample. Therefore, it enables the determination of substrate specificities and

excision kinetics of DNA glycosylases by identifying which lesions are or are not excised

from DNA by a given DNA glycosylase. Subsequently, substrate specificities and excision

kinetics of numerous DNA glycosylases has extensively been studied (reviewed in

[22,216,217]).

20

3.1.1. Substrate specificities of prokaryotic DNA glycosylases

The concept described above has been used for the first time to investigate the substrate

specificity of E. coli Fpg [41]. This enzyme had originally been shown to recognize and remove purine-derived lesions with an opened imidazole ring such as 2,6-diamino-4- hydroxy-N7-methyl-5-formamidopyrimidine (Me-FapyGua) (derived from 7-methylguanine) and FapyAde [218-220]. Subsequent work reported the excision by E. coli Fpg of 8-OH-Gua

[221], and pyrimidine-derived lesions 5-OH-Cyt and 5-OH-Ura from oligodeoxynucleotides

[222]. The use of GC-MS demonstrated the excision of 8-OH-Gua, FapyGua and FapyAde by this enzyme, but no excision of pyrimidine-derived lesions from damaged DNA containing

multiple lesions [41] (for the structures of these compounds see Figure 1). A subsequent

study reported the excision of these three lesions by similar Michaelis-Menten kinetics [223].

These results clearly showed that FapyGua and FapyAde may also be the main substrates of

E. coli Fpg in cells, in contrast to the claim of 8-OH-Gua being the main physiological

substrate of this enzyme without providing any data on the former two compounds [224,225].

Five mutant forms of E. coli Fpg have been generated and used to investigate the effect of

single point mutations in the fpg gene targeting highly conserved amino acids on the

specificity of this enzyme [226]. The results showed that single mutations targeting amino

acids Lys-57, Lys-155 and Pro-2 dramatically affected the specificity up to a complete loss of

activity. A protein homologous to E. coli Fpg has been isolated from the bacterium

Deinococcus radiodurans that exhibits resistance to the effects of extreme doses of ionizing

radiation and other DNA-damaging agents [227,228]. This enzyme designated Deinococcus

radiodurans Fpg efficiently excised 8-OH-Gua, FapyGua and FapyAde similar to E. coli Fpg,

but with significantly different excision kinetics [229].

E. coli Nth of the Nth superfamily exhibits a broad substrate specificity for cytosine- and

thymine-derived lesions [222,230-234]. The use of GC-MS extended the substrate specificity

21

E. coli Nth for pyrimidine-derived lesions and also included purine-derived FapyAde

[235,236]. Most of the pyrimidine-derived lesions listed in Figure 1 have ben found to be the

substrates of this enzyme. Another DNA glycosylase endonuclease VIII (Nei) of E. coli exhibits strong homology to E. coli Fpg and other bacterial Fpg proteins, but no significant sequence similarity to E. coli Nth [237-240]. Both enzymes have overlapping substrate specificity [222,234]. The use of the GC-MS and damaged DNA with multiple lesions extended the substrate specificity of E. coli Nei and showed that this enzyme also excises

FapyAde as E. coli Nth does [241]. E. coli uracil DNA glycosylase (UNG), which is specific for removal of uracil from DNA [242], has also been found to act on cytosine-derived products 5-OH-Ura and isodialuric acid (5,6-dihydroxyuracil) [243,244]. E. coli MutY of the

Nth superfamily removes adenine paired with 8-OH-Gua [245].

3.1.2. Substrate specificities of eukaryotic DNA glycosylases

Structural and functional homologues of E. coli Nth have been found in yeast and humans

[246,247]. Schizosaccharomyces pombe Nth (S. pombe Nth) efficiently excised 5-OH-Cyt, 5-

OH-Ura, Thy glycol, 5-OH-6-HThy and 5,6-diOH-Cyt from DNA, exhibiting a narrower substrate specificity than E. coli Nth [246,248]. Human NTH1 acted on the same DNA lesions as S. pombe Nth, albeit with significant differences in excision kinetics [249].

Efficient excision of purine-derived FapyAde by NTH1 has also been observed [250].

FapyAde accumulated in nth1–/– mice, providing the evidence that FapyAde is the

physiological substrate of NTH1 [250,251]. (Fig. 4) illustrates the levels of FapyAde and 8-

OH-Gua in livers of nth1–/–, ogg1–/– and ogg1–/–/nth1–/– mice, demonstrating the accumulation

of FapyAde in nth1–/– mice when compared to wt mice, but not in ogg1–/– mice, and that of 8-

OH-Gua in ogg1–/–/mice, but not in nth1–/– mice. As expected, both compounds accumulated

in double knockout animals. The use of liver mitochondrial and nuclear extracts of wild type

(wt)-mice and nth1‒/‒, ogg1–/– and ogg1–/–/nth1–/– mice, and oligodeoxynucleotides containing

22

FapyAde confirmed this finding and also that FapyAde is not a substrate of OGG1 [250], as

had been previously shown in in vitro experiments [252]. The repair of 5,6-dihydrouracil has

also been found to be reduced, but not nullified in nth1–/– mice [253].

Schizosaccharomyces cerevisia possesses two DNA glycosylases/AP lyases named Ntg1

and Ntg2, which relate to each other and to their functional homologue E. coli Nth [254-256].

These DNA glycosylases have been shown to excise a number of pyrimidine-derived lesions, and purine-derived lesions FapyAde and FapyGua [257,258]. Thus, the cross-activities of

Ntg1 and Ntg2 are clearly different from E. coli Nth and Nei in that they efficiently remove

FapyGua in addition to FapyAde. Human uracil DNA glycosylase, which is mainly specific for removal of uracil from DNA [259], has been found to also excise cytosine-derived

products isodialuric acid, 5-OH-Ura and alloxan from DNA containing multiple lesions

[260,261]. Similarly, human SMUG1 exhibited specificity for the same lesions [261].

Functional homologues of E. coli Fpg, named OGG1, have been discovered in eukaryotes

[262-266]. All these DNA glycosylases exhibited an identical substrate specificity with the

excision of 8-OH-Gua and FapyGua from damaged DNA with multiple lesions, although excision kinetics somewhat varied among the enzymes [252,264,267-269]. The failure of

FapyAde excision by OGG1 from eukaryotes indicates significant differences between these enzymes and E. coli Fpg. Two different types of human OGG1 (hOGG1) have been discovered and designated α-hOGG1 and β-hOGG1 [263]. α-hOGG1 is targeted to the nucleus, whereas β-hOGG1 is located in the mitochondrion [270,271]. Two forms of α-

hOGG1 due to a polymorphism at codon 326, α-hOGG1-Ser326 and α-hOGG1-Cys326, have

been found in human cells [272,273]. α-hOGG1-Ser326 and α-hOGG1-Cys326 efficiently excised FapyGua and 8-OH-Gua from damaged DNA with multiple lesions, with the former exhibiting ≈ 2-fold greater activity than the latter [252]. Both forms had a greater preference for FapyGua than 8-OH-Gua. Two mutated forms of α-hOGG1-Ser326, α-hOGG1-Gln46 and

23

α-hOGG1-His154, have been found in tumor cells [268,274,275]. Both forms exhibited

efficient activity on FapyGua and 8-OH-Gua; however, their activity was lower than that of

α-hOGG1-Ser326 [268]. MUTYH, which is a homologue of E. coli MutY, also plays a role in the repair of 8-OH-Gua by removing adenine paired with it [193,276,277].

E. coli Nei-like DNA glycosylases have been discovered in eukaryotes, and named

NEIL1, NEIL2 and NEIL3 [278-285]. NEIL1 is regulated and may thus be associated with the replication fork [286-289]. It is located to both the nucleus and mitochondrion, lending credit to its importance in maintaining the genetic stability [250].

NEIL1 mainly removes FapyAde and FapyGua from damaged DNA with multiple lesions,

and also Thy glycol and 5-OH-5-MeHyd albeit to a lesser extent [278,290]. In agreement

with in vitro studies, FapyAde and FapyGua have been shown to be the main physiological

substrates of NEIL1 in vivo [251,291]. NEIL1 exhibits no detectable activity toward 8-OH-

Gua in vitro or in vivo [122,278,290-293]. Recently, an additional specificity of NEIL1 has

been discovered. The accumulation of R-cdA and S-cdA in liver DNA of neil1‒/‒ mice has

been observed [122]. Since R-cdA and S-cdA are repaired by NER and not by BER [108,294-

297,297,298], this finding suggested that NEIL1 may be involved in NER in addition to its function as a DNA glycosylase in BER. To this end, there is evidence that NEIL1 may

interact with proteins of the NER complex. For example, similar to NEIL1, the Cockayne

syndrome complementation group B protein (CSB) plays a role in the repair of S-cdA [120],

and stimulates the action of NEIL1 on FapyAde and FapyGua [291]. This is in contrast to its lack of interaction with OGG1 [299]. Moreover, CSB and NEIL1 coimmunoprecipitate and

colocalize in HeLa cells, indicating that these proteins cooperate in the repair of

formamidopyrimidines [291]. NEIL1 may also interact with other NER proteins and

accelerate the repair of R-cdA and S-cdA, since it cannot itself initiate the BER at these

lesions.

24

The other Nei-like DNA glycosylase NEIL2 has been shown to preferentially excise pyrimidine-derived lesions from oligodeoxynucleotides with bubble structures [280,300]. It also exhibited specificity for removal of Sp and Gh from oligodeoxynucleotides [301]. Thus far, however, no excision by NEIL2 of any base lesions from DNA with multiple lesions has been observed. FapyAde and FapyGua have also been found to be the main substrates of

NEIL3; however, the efficient removal by this enzyme of pyrimidine-derived lesions has been observed from DNA with multiple lesions as well [293]. 8-OH-Gua has been shown to be not a substrate of NEIL3, either; nevertheless, its oxidation products Sp and Gh were efficiently removed by NEIL3 and also by NEIL1 from synthetic oligodeoxynucleotides

[301,302].

3.1.3. Repair of sugar lesions

Sugar lesions within DNA or bound to DNA as end groups (Figure 2), the oxidatively induced formation of which was discussed above, constitute AP sites with a modified 2'- deoxyribose moiety in DNA. On the other hand, AP sites with the intact 2'-deoxyribose moiety are formed in DNA of living cells by spontaneous hydrolysis of the glycosidic bond that occurs several thousand times per day per cell and also by the action of DNA glycosylases on sites with damaged DNA bases (reviewed in [11,303]). BER is the primary pathway that repairs AP sites in mammalian cells. Although multifunctional DNA glycosylases can cleave AP sites, apurinic/apyrimidinic endonuclease 1 (APE1) processes most AP sites via hydrolysis of the phosphate bond 5' to the AP site creating a single strand break with a 3'-OH group and a 5' terminal 2'-deoxyribose phosphate (dRP) residue

[304,305]. Subsequently, DNA polymerase β (Pol β) and other enzymes repair the remaining nick. In certain circumstances, AP sites are also repaired by BER via long-patch repair [306].

2-Deoxypentose-4-ulose, erythrose and 2-deoxyribonic acid (or its lactone form) within

DNA, and 2-deoxytetradialdose as an end group (Figure 2) are the substrates of the first step

25 of BER involving APE1, Pol β and Pol λ; however, they are not processed as efficiently as

AP sites with the intact 2'-deoxyribose moiety [307-319]. 2-Deoxyribonic acid is repaired

almost exclusively by long-patch BER [315,320-322]. The sugar lesions can cause

irreversible inhibition of BER enzymes and the formation of interstrand DNA cross-links

[317,319,323-333]. Furthermore, DNA-protein cross-links occur between 2-deoxyribonic acid and the proteins E. coli endonuclease III, Pol β and histones [310,311,315,321,334,335].

Thus, the action of APE1 in short-patch BER on 2-deoxyribonic acid is stalled by the formation DNA-protein cross-links between this lesion and Pol β [315]. On the other hand, long-patch BER prevents the formation of DNA-protein cross-links between 2-deoxyribonic acid and Pol β [320].

3.2. Nucleotide excision repair

Nucleotide excision repair removes bulky DNA-distorting lesions from DNA [209,336-

341]. Global genome repair and -coupled repair constitute two distinct mechanisms of NER and are responsible for the repair of the entire genome and preferential repair of transcribing DNA strands, respectively. NER has also been reported to repair oxidatively induced lesions such as thymine glycol and 8-OH-Gua [342-344]. Almost three decades ago, it has been proposed that 8,5'-cyclopurine-2'-deoxynucleosides would not be repaired by BER because of the 8,5'-covalent bond between the base and sugar moieties, and thus they would likely be subject to repair by NER [108,294]. Indeed, NER, not BER, has been shown to repair R-cdA and S-cdA, with the former being repaired more efficiently than the latter [295,296]. On the other hand, inefficient repair of R-cdG and S-cdG has been reported by E. coli NER enzymes in vitro [345]. Recently, a number of DNA glycosylases including Fpg, NEIL1 and OGG1 have been tested on oligodeoxynucleotides containing cdA or cdG; however, no cleavage has been detected, confirming the lack of activity of BER on

26

8,5'-cyclopurine-2'-deoxynucleosides [297]. Furthermore, these DNA glycosylases at high

concentrations failed to form DNA-protein complexes with oligodeoxynucleotides containing

S-cdA or S-cdG. In contrast, HeLa cell extracts excised 24-32 base-pair fragments from long

double-stranded oligodeoxynucleotides containing S-cdG or S-cdA, with a greater efficiency

for the former than the latter [297]. The efficiency of repair depended on the complementary

base opposite the lesion. Just recently, the structural basis for the recognition of 8,5'-

cyclopurine-2'-deoxynucleosides by NER has been investigated in detail [298]. Using

extracts of HeLa cells and oligodeoxynucleotides containing R-cdA, S-cdA, R-cdG or S-cdG with identical sequence contexts, NER has been shown to excise the R-diastereomers of both cdA and cdG with a ≈ 2-fold greater efficiency than their S-diastereomers. However, the overall excision efficiencies between cdA and cdG were similar. The R-diastereomers of cdA and cdG caused greater distortion of the DNA backbone than their S-diastereomers, correlating with NER incision efficiencies. Recently, the apurinic/apyrimidinic endonucleases, E. coli Xth and human APE1 have been reported to remove S-cdA at the 3' terminus of duplex DNA, but not that located at 1 or more away from this end

[346]. This mechanism has been suggested as a complementary pathway to NER to remove

S-cdA and possibly other 8,5'-cyclo-2'-deoxynucleosides as well. This is an intriguing mechanism; however, the formation of an 8,5'-cyclo-2'-deoxynucleoside as the end unit of a broken DNA strand must occur for this mechanism to be active. In support of this notion, evidence has recently been provided for the incorporation of R- and S-diastereomers of cdATP into DNA by replicative DNA polymerases, inhibiting further DNA synthesis and thus generating a DNA strand with a cdA at the 3'-terminus [347].

27

4. Genetic effects of oxidatively induced DNA lesions

Failure to repair DNA lesions before replication of DNA-damaged cells may lead to cell death, cytotoxicity and mutagenicity, and ultimately to disease processes including carcinogenesis. If a DNA lesion is not removed from DNA, it may be tolerated and bypassed by DNA polymerases, which may mispair it with a non-cognate intact DNA base, leading to a following the next step of replication. On the other hand, a DNA lesion can block the action of DNA polymerases to perform DNA synthesis, thus becoming a lethal lesion leading to cell death. The interplay between the repair and the two types of replication may ultimately determine the future of a DNA-damaged cell and potentially that of the organism.

It should be mentioned that a DNA lesion can also pair with a cognate DNA base, which will make it neither lethal nor mutagenic [348].

4.1. Purine -derived lesions

Among the oxidatively induced DNA lesions, 8-OH-Gua has been the most investigated lesion in terms of its biological effects, perhaps at the expense of the other equally important lesions. In an early work, this lesion has been shown to induce numerous mutations in vitro that included a G → T transversion mutation indicating mispairing of 8-OH-Gua with Ade

[349]. In a subsequent work, the genetic effects of 8-OH-Gua have been determined after transfection of the single-stranded site-specifically modified viral genome into wild-type E. coli [350]. This work provided direct evidence that 8-OH-Gua is premutagenic in vivo and leads to the G → T transversion mutation as a major mutagenic event. The bypass efficiency of 8-OH-Gua generally amounted to 85-90% [351]. Frequency of mispairing depended on the polymerase [352,353]. 8-OH-Gua has also been shown to pair with cognate Cyt, albeit to a lesser extent, causing no mutations [350,352,354-356]. Fig. 5 illustrates the 8-OH-Gua•Ade mispair [356]. Oxidation products of 8-OH-Gua such as Sp and Gh have been found to

28 exhibit mutagenic effects as well as cytotoxic effects depending on the polymerase, sequence context, etc. (reviewed in [72]). The other equally important guanine-derived lesion FapyGua has been shown to also mispair with non-cognate Ade (Fig. 5), leading to G → T transversion mutations [357,358]. In simian kidney cells, FapyGua has even been more mutagenic than 8-

OH-Gua [359,360], although a weak mutagenicity for FapyGua in E. coli has been observed

[361]. A subsequent work unequivocally demonstrated significant in vivo mutagenicity of

FapyGua in an E. coli triple mutant lacking Fpg, Nei and MutY glycosylase activities by expressing glycosylase domains of mouse NEIL1 and NEIL3 in these cells [293]. G → T transversion mutations are the second most common somatic mutations found in human cancers, constituting 14.6% of all mutations in the tumor suppressor gene TP53 [362]. Of course, this does not mean that these mutations entirely result from 8-OH-Gua and FapyGua.

Other DNA lesions may lead to such mutations as well.

8-OH-Ade has been shown to pair with cognate Thy and to mispair with Gua and Ade depending on the polymerase [363-366]. Fig. 5 illustrates the 8-OH-Ade•Gua mispair [367].

The lack of mutagenicity of 8-OH-Ade in bacteria has been reported [368]. On the other, this lesion caused A → G transition and A → C transversion mutations in mammalian cells at a mutation frequency of ≈1% only [366,369]. FapyAde, which has the precursor C8-OH‒ adduct radical of Ade in common with 8-OH-Ade [10], has been found to direct Klenow exo‒

fragment to misincorporate Ade opposite itself, potentially leading to A → T transversions

[370]. The FapyAde•Ade mispair is illustrated in Fig. 5 [370]. Both 8-OH-Ade and FapyAde

were very weakly mutagenic in simian kidney cells when compared to 8-OH-Gua and

FapyGua [359]. In the case of another adenine-derived product, 2-OH-Ade, DNA polymerases inserted all DNA bases opposite this lesion with the possibility of leading to all mutations involving Ade [371,372]. In E. coli, the frequency and the spectrum of the mutations depended on the sequence contexts and the lagging/leadings template strands

29

[373]. In simian COS-7 cells, the primary mutation has been a ‒1 deletion followed by A →

G transitions and A → T transversions [374]. The observed mutation frequencies (0.6 % ‒ 0.1

%) have been comparable with those of 8-OH-Gua in NIH3T3 cells [375,376]. The results

suggested the formation of 2-OH-Ade•Cyt and 2-OH-Ade•Ade mispairs in living cells. In

addition, 2-OH-Ade formed in the nucleotide pool induced G•C → T•A transversions in E.

coli [377]. Moreover, 2-OH-dATP led to G•C → A•T transitions and G•C → T•A

transversions during in vitro replication using HeLa cell extracts, albeit to a lesser extent

[378].

4.2. Pyrimidine-derived lesions

Thymine glycol is one of the major and most investigated products of thymine. Its biological effects have also been investigated extensively. Thymine glycol correctly pairs with Ade and is thus poorly mutagenic [348,379-382]. In some sequence contexts, however, it is bypassed by DNA glycosylases and can pair with non-cognate Gua to a low extent, leading to T → C transitions [383-386]. In general, Thy glycol constitutes a strong block to

DNA polymerases and is thus a lethal lesion [348,379,382,383,386-391]. 5-Hydroxy-6- hydrothymine also strongly blocks an E. coli DNA polymerase and is a lethal lesion [392].

5,6-Dihydrothymine pairs with cognate Ade and is not a block to DNA polymerases; therefore, it is neither lethal nor mutagenic [393,394].

Deamination and dehydration reactions of cytosine glycol result in the formation of 5-OH-

Cyt, 5-OH-Ura and uracil glycol [91,235]. These lesions can exist in DNA simultaneously

[235]. Uracil glycol, 5-OH-Cyt and 5-OH-Ura pair with non-cognate Ade leading to C → T transitions [348,380,386,395-397]. Fig. 6 illustrates of the pathway of the formation of C →

T transitions due to mispairing of 5-OH-Ura and 5-OH-Cyt with Ade. The incorporation of the intact base opposite these lesions depended on the sequence context. Thus, C → G

30 transversions have also been observed as a result of the mispairing of 5-OH-Cyt with Cyt

[398]. The minor anionic imino tautomer of 5-OH-Cyt has been shown to mispair with Ade

(Fig. 5) and to be the likely source of C → T transitions caused by this lesion [399]. In E. coli, 5-OH-Ura and uracil glycol led to C → T transitions with high frequency (83% and

80%, respectively), whereas 5-OH-Cyt elicited the same mutations with a lower frequency

(0.05%) and C → G transversions even to a much lesser extent [397]. These results suggested

that 5-OH-Ura and uracil glycol may be the main cause of C → T transitions in cells. These

transition mutations have been found to be the most frequently occurring mutations in human tumors and in TP53 [362,400], and from oxidatively induced DNA damage [396,401].

However, this does not mean that 5-OH-Ura and uracil glycol would be the only source of C

→ T mutations. For example, errors by replicative DNA polymerases [402,403], and deamination of 5-methylcytosine can also cause these mutations at high frequency [15,404].

4.3. Sugar lesions

AP sites with an intact 2'-deoxyribose are mutagenic [405,406]. Similarly, the sugar lesions 2-deoxypentose-4-ulose, erythrose and 2-deoxyribonic acid exhibit mutagenic effects

[322,407-412]. They also form interstrand DNA cross-links and DNA-protein cross-links, and thus are converted to other types of DNA damage such as DNA strand breaks that may additionally be deleterious to living cells [310,311,315,319,321,329-334,413]. The ability of

2-deoxypentose-4-ulose and 2-deoxytetradialdose to inactivate Pol β and Pol λ indicates that these lesions may be a significant source of the cytotoxicity caused by DNA-damaging agents

[317,326-328]. Furthermore, there is evidence that strand scission at AP sites and at the sugar lesions 2-deoxypentose-4-ulose and 2-deoxyribonic acid are accelerated in nucleosome core particles [319,414-418]. In early studies, the treatment of DNA with neocarzinostatin chromophore (NCS) resulted in significant amounts of G•C → A•T transitions, pointing to 2-

31 deoxyribonic acid as the highly likely source of these mutations [419,420]. Indeed, 2- deoxyribonic acid has been identified at the site of NCS-induced Cyt release in a certain sequence [421,422], and confirmed to be a source of transition mutations [407,409].

Incorporation of Ade by Klenow opposite this lesion is on a par with the observed mutations

[423]. In addition, 2-deoxyribonic acid constitutes a stronger block to the translesional synthesis by Klenow than the AP site with an intact 2'-deoxyribose [423,424]. Cross-linking

occurring between 2-deoxyribonic acid and histones may also contribute to genotoxic effects

of this lesion in DNA [315,335]. Similar to 2-deoxyribonic acid, Ade is preferentially

incorporated opposite erythrose within DNA, which exhibits an effect on Klenow exo‒

similar to that of the AP site with an intact 2'-deoxyribose moiety [313]. Thus, erythrose may be a mutagenic lesion as well. Gua was also incorporated, albeit to a much lesser extent, but not Cyt or Thy to any significant extent. On the other hand, Pol V in E. coli preferentially incorporated Gua opposite 2-deoxyribonic acid, even more than opposite the AP site, and discriminated between Gua and Ade [425]. Moreover, 2-deoxyribonic acid induced significant Gua incorporation in E. coli under SOS conditions [411]. In the case of 2-

deoxypentose-4-ulose, Klenow exo‒ efficiently incorporated Ade opposite, followed by Gua

with ≈ 10-fold greater preference for the former than the latter [426]. Klenow exo+ exhibited

slightly less preference for Ade incorporation than Klenow exo‒. Although the sugar lesions

and AP sites with an intact 2'-deoxyribose moiety have substantial structural differences, they

exhibit similar effects on the activities of Klenow exo+/exo‒ fragments and bypass polymerases. Taken together, sugar lesions in DNA (Fig. 2) apparently exhibit significant mutagenic and genotoxic effects that rival those exhibited by modified DNA bases.

32

4.4. Tandem lesions

4.4.1. 8,5'-Cyclopurine-2'-deoxynucleosides

Genetic effects of 8,5'-cyclopurine-2'-deoxynucleosides have drawn much attention in the past decade and have been investigated quite extensively. S-cdA blocks transcription and several DNA polymerases, reduces transcription and causes transcriptional mutagenesis

[295,296,427-429]. Multiple nucleotide deletions occur due to incorporation of adenosine opposite to the next 5'- to S-cdA by RNA polymerase II [429]. Recently, S-cdG has been found to be a strong block to replication and a potent DNA polymerase V-dependent mutagenic lesion in E. coli, leading mainly to G → A transitions, indicating it mispairs with

Thy, and to G → T transversions to a lesser extent [345]. Fig. 5 illustrates the S-cdG•Thy mispair [297,430]. The pathway leading to G → A transitions by this mispair is shown in Fig.

6. Similarly, S-cdG and S-cdA strongly block DNA replication, and also cause G → A transitions and A → T transversions in five different strains of E. coli to an extent of 20% and

11%, respectively [431]. On the other hand, both human polymerase η and S. cerevisiae polymerase η bypassed S-cdG and S-cdA accurately and efficiently, indicating that mutagenic bypass of these lesions may relate to other DNA polymerases [432]. Indeed, S-cdG and S- cdA have been shown to be strong blocks to replication by DNA polymerase IV, exo-free

Klenov fragment and Dpo4 [433]. Furthermore, the same study showed the occurrence of both A → T transversions and A → G transitions caused by S-cdA in equal frequency in wild-type E. coli; however, the results with DNA polymerase IV-deficient strain suggested that DNA polymerase IV had played a role in A → G transitions caused by S-cdA. DNA polymerase IV incorporated Cyt and Thy opposite S-cdA with almost equal efficiency, whereas the cognate Cyt was efficiently inserted opposite S-cdG by this polymerase. On the other hand, the incorporation of Thy by the exo-free Klenov fragment was more than twice

33 that of Cyt by DNA polymerase IV [433]. The S-cdA•Thy pair (Fig. 5) has been found to be

very stable when compared to the Ade•Thy pair [434].

4.4.2. Tandem lesions

Two tandem lesions identified in vivo, i.e., Gua[8,5]Cyt and Gua [8,5-Me]Thy (Fig. 3),

have been shown to be cytotoxic and mutagenic [151,159,160]. In E. coli, Gua[8,5]Cyt blocked DNA replication and exhibited significant mutagenicity in vivo due to misincorporation opposite the 5'-guanine moiety that included G → T and, to a lesser extent,

G → C transversion mutations. DNA polymerase V was responsible for the mutagenicity of this lesion. S. cerevisiae DNA polymerase η inserted the cognate dGMP opposite 3'-cytosine moiety, indicating that the 3'-cytosine moiety of this lesion is not mutagenic. However, this polymerase inserted non-cognate dAMP and dGMP opposite the 5'-guanine moiety, which may result in G → T and G → C transversion mutations. The Klenow fragment of E. coli

DNA polymerase I stopped synthesis after incorporating the cognate dAMP opposite 3'- thymine moiety of Gua[8,5-Me]Thy; however, S. cerevisiae DNA polymerase η performed synthesis past this lesion and incorporated non-cognate dAMP and dGTP opposite 5'-guanine.

Thus, Gua[8,5-Me]Thy may give rise to G → T and G → C transversion mutations.

Gua[8,5]Cyt and Gua[8,5-Me]Thy may distort the DNA structure, leading to lack of recognition by DNA polymerases of the hydrogen bonding property of the 5'-guanine moiety that assumes a syn configuration in contrast to the anti configuration of the 3'-cytosine moiety

[151]. Thus, purine nucleotides may be inserted opposite 5'-guanine more efficiently than pyrimidine nucleotides. Mutagenic effects of the Thy-Tyr cross-link have not been investigated.

34

5. Oxidatively induced DNA damage and cancer

DNA repair is essential to life. Thus, unrepaired DNA lesions may lead to detrimental consequences in living organisms. DNA lesions that escape repair may accumulate in the genome, progressively leading to increase in mutation rate, i.e., mutator phenotype, and thus in genetic instability, a hallmark of cancer [3,12,14,15,191,435,436]. Defective DNA repair is associated with carcinogenesis [14,15,19,437-455]. Persistent oxidative stress exists in cancer

[456]. In agreement with this fact, precancerous and cancerous tissues or cancer cell lines have been shown to contain oxidatively induced DNA lesions at elevated levels when compared to surrounding cancer-free tissues or to normal cell lines [112,113,116,117,457-

470]. Most of these lesions are mutagenic (see above) and may thus play a significant role in carcinogenesis and other disease processes. DNA repair genes are also prone to mutations.

Germline mutations cause genetic instability and predisposition to cancer [191,449].

Polymorphisms of DNA repair genes including the BER genes also increase the cancer risk

and may determine the outcome of disease for patients [191,444,449,454,471-477]. Therapy

resistance occurs in tumors and may be due to defects in DNA repair, adversely affecting the

outcome of patient survival [16-18,449,454,478-480]. Several studies reported lower levels of

ethano-DNA adducts and oxidatively induced DNA lesions in cancerous tissues than in

surrounding non-cancerous tissues, providing the evidence that DNA repair may be

upregulated in cancerous tissues [123,481,482]. Upregulated DNA repair in cancerous tissues

may cause resistance to therapeutic agents. Despite their adverse effects, DNA repair

alterations may have the potential to help develop the concept of personalized cancer therapy

and may also serve as promising predictive cancer biomarkers [480].

35

5.1. Role of DNA glycosylases of BER in carcinogenesis

As mentioned above, most of oxidatively induced DNA base lesions, except for 8,5'-cyclo-

2'-deoxynucleosides which are subject to NER, are repaired by BER with the action of

various DNA glycosylases in the first step of this mechanism. BER removes numerous

endogenously and exogenously induced lesions from cellular DNA per day and thus protects

the genetic stability and plays an important role in disease prevention including cancer

prevention. Thus, defects in BER are associated with neurological disorders and cancer

[214,476,477,483]. Polymorphic variants of DNA glycosylases have been found in human

populations in connection to various cancer incidences. Moreover, mouse models with

knockouts of DNA glycosylase genes have been developed to study resulting phenotypes.

5.1.1. OGG1

OGG1 has been one of the most investigated DNA glycosylases. Among many oxidatively induced DNA lesions (Fig. 1), 8-OH-Gua and FapyGua are the main physiological substrates of human OGG1 (hOGG1) and its eukaryotic homologues. This fact has been proven in

experiments in vitro and in vivo (discussed above). When compared to wt-mice, simultaneous accumulation of both lesions has been demonstrated in livers of ogg1‒/‒ mice; however, no

accumulation of FapyAde has been observed [122,250]. These observations are on a par with the in vitro findings of the substrate specificity of OGG1 (see above). These results have also been confirmed using liver mitochondrial and nuclear extracts of wt-mice and ogg1‒/‒ mice,

and oligodeoxynucleotides containing FapyAde, FapyGua or 8-OH-Gua at a defined position

[250]. Human ogg1 gene contains a variety of single-nucleotide polymorphisms

(http://www.ncbi.nlm.nih.gov/sites/entrez?db=snp) [273,275,452,484-486]. The most

common polymorphic variant of hOGG1 is OGG1-Cys326 with a high frequency in human population [452]. Many studies provided the evidence for the association of this variant with

36 the risk of a number of cancers [272,274,275,447,452,484,487-511]. On the other hand, no risk of some cancers with OGG1-Cys326 has been demonstrated [512-514]. Other polymorphic variants of hOGG1 such as OGG1-His154, OGG1-Gln46, OGG1-Gln209,

OGG1-Thr321, OGG1-His154, OGG1-131Gln have also been found in some cancer cell lines and tissues [268,272,274,275,515,516]. Differences between the activities of the variants of

OGG1 and the wt-OGG1 have been observed. OGG1-Cys326, OGG1-His154, OGG1-Gln46 and OGG1-Asn322 exhibited significantly lower activities than that of wt-OGG1

[252,268,517]. In contrast, excision kinetics of one variant, i.e., OGG1-Val288 has been

found to be similar to that of wt-OGG1 [517]. Low OGG1 activity has been shown to

constitute a risk factor in some types of cancers [447,518-521]. The expression of OGG1 in

eighteen human cancer and three normal cell lines has been studied [468]. Sixteen of the

cancer cell lines exhibited overexpression of OGG1; however, two of them had even lower

expression of OGG1 than normal cell lines. Increased levels of 8-OH-Gua have been

observed in these two cancer cell lines. Their mitochondria also exhibited compromised

repair of 8-OH-Gua, suggesting OGG1 plays a role in the maintenance of the mitochondrial

genome. Overall, the results implicated 8-OH-Gua repair defects in certain lung cancers.

Despite all these findings discussed above, ogg1‒/‒ mice have been found to be viable with no tumor formation, although they exhibited organ-specific accumulation of 8-OH-Gua and

elevated G → T transversion mutations [522-524]. Treatment of such mice with KBrO3

caused a dramatic increase in the level of 8-OH-Gua in both livers and kidneys; however, no

tumor formation has been observed in these organs [525-527]. Furthermore, exposure of both

wt-mice and ogg1–/– mice to low doses of ionizing radiation, ogg1–/– mice exhibited a

significant increase in G → T transversions in their brains; however, no tumor development

has been observed [528]. On the other hand, an increase in spontaneous lung adenomas and

carcinomas along with a significant accumulation of 8-OH-Gua has been observed in a

37 different strain of ogg1‒/‒ mice [529]. However, 8-OH-Gua has been the only measured lesion

among many other DNA lesions (Fig. 1) to unequivocally conclude that this effect had been due to its accumulation only. Additional knockout of the mth1 gene in these mice further increased the level of 8-OH-Gua with no significant increase in carcinogenesis. Exposure to chronic UVB radiation increased the level of 8-OH-Gua in ogg1–/– mice and to make them susceptible to skin carcinogenesis [530]. Again, no other lesion among many DNA lesions has been measured to draw useful conclusions about the role of 8-OH-Gua in the observed carcinogenesis. Due to all these facts discussed above, there seems to be no clear association of ogg1 deficiencies and the accumulation of 8-OH-Gua with human carcinogenesis. On the other hand, mice with deficiencies in both ogg1 and mutyh exhibited the incidence of tumors predominantly in lung and ovarian tumors, and lymphomas with accompanying G → T transversions in the majority of lung tumors at codon 12 of the K-ras oncogene [531]. Mutyh–

/– mice developed no tumors similar to ogg1–/– mice. In another study using the same strain of

mutyh–/– mice, no age-associated accumulation of 8-OH-Gua has been observed in several organs except for liver; however, mutyh–/–/ogg1–/– mice accumulated 8-OH-Gua and exhibited

the incidence of lung and small intestine cancers [532]. These findings may indicate the

requirement for carcinogenesis of synergistic occurrence of mutations in several DNA repair

genes.

5.1.2. NEIL proteins

Extensive studies in the past decade demonstrated a critical role for NEIL1 in the

maintenance of the genetic stability and disease prevention. Polymorphic variants of NEIL1

have been discovered in humans. Three NEIL1 variants, NEIL-Arg242, NEIL1-Arg245 and

NEIL1-Gly334 have been found in human gastric cancers; however, they exhibited activity similar to that of wt-NEIL1 [533]. A NEIL1 deletion variant with no activity has also been observed in gastric cancers [533]. Three promoter polymorphisms in neil1 have been

38 identified in gastric cancers, although the consequences of these mutations have not been

determined [534]. Accumulation of mutations in the hprt locus and their increase by

oxidative stress have been observed in neil1-knockdown human bronchial cells and Chinese

hamster ovary cells [535]. Knocking down neil1 rendered embryonic stem cells more sensitive to killing effects of ionizing radiation [536]. Exposure of human carcinoma cells to

oxidative stress increased expression of NEIL1 [537]. NEIL1-Asp83 and NEIL1-Lys181

variants have been found in patients with cholagiocarcinoma and primary sclerosing

cholangitis, respectively [538]. Rare NEIL1-Ser203 and NEIL1-Gln339 variants have been

observed in patients with colorectal adenomas; however, the latter variant also existed in a

control individual [516]. Four polymorphisms have been reported for the human neil1 [292].

Corresponding variants NEIL1-Cys82, NEIL1-Asp83, NEIL1-Asn252 and NEIL1-Arg136 have been isolated and characterized [292]. An AP site- or a Thy glycol-containing oligodeoxynucleotide, and DNA substrates with multiple lesions have been used to test the

AP-lyase and glycosylase activities of these variants in comparison to wt-NEIL1. NEIL1-

Cys82 and NEIL1-Asn252 exhibited a β,δ-elimination activity on the AP-site as wt-NEIL1.

However, NEIL1-Asp83 had a β-elimination activity only and NEIL1-Arg136 exhibited no activity at all. An efficient excision of FapyAde and FapyGua from DNA with multiple lesions and that of Thy glycol from an oligodeoxynucleotide by wt-NEIL1, NEIL1-Cys82 and NEIL1-Asn252 has been observed. In contrast, no DNA glycosylase activity whatsoever has been detected for NEIL1-Asp83 and NEIL1-Arg136. The measurement of the specificity constants revealed a preference of NEIL1 and its active variants for FapyAde over FapyGua, perhaps because FapyGua is also a substrate of OGG1, which does not act on FapyAde, as was discussed above. Overall, this work suggested that individuals with neil1 mutations may be at risk for disease development.

39

A neil1–/– mouse model has been developed to investigate the consequences of NEIL1

deficiency in vivo [251,539-541]. Male neil1–/– mice developed obesity by 24 months of age,

accompanied by severe fatty liver, increased circulating lipids and hyperinsulinemia, which

are collectively called metabolic syndrome. Female neil1–/– mice also developed obesity, but

to a lesser extent than male neil1–/– mice [539]. In this context, there is evidence that

metabolic syndrome may be associated with certain types of cancer [542-545]. When exposed

to oxidative stress, male neil1–/– mice gained more weight than wt-mice [539,541]. Mutations

increased in mitochondrial (mt) DNA and the mtDNA content decreased in livers of these

mice. All these results strongly suggested that NEIL1 may play an important role in disease

prevention and that NEIL1 deficiency may lower the threshold for tolerance of oxidatively

induced DNA damage.

In another study, nth1–/– and neil1–/–/nth1–/– mice in addition to neil1–/– mice have been developed to understand the roles of NEIL1 and NTH1 in carcinogenesis [251]. Very few tumors have been observed in the first year of these animals’ lives. During the second year, however, neil1–/– and nth1–/– male and female mice developed pulmonary tumors and

hepatocellular tumors to a low extent up to ≈ 4 % and ≈ 15%, respectively. However, nth1–/–

/neil1–/ – mice exhibited a dramatic increase in both cancer incidences at a much greater rate

(up to ≈ 75 %) than neil1–/– or nth1–/– animals [251]. These results are illustrated in Fig. 7.

Pulmonary tumors contained activating GGT → GAT transitions in codon 12 of their K-ras

oncogene. This is in contrast to the activating GGT → GTT transversions in codon 12 of the

K-ras oncogene of the pathologically similar pulmonary tumors in ogg1–/–/mutyh1–/– mice

[531]. Oxidatively induced DNA lesions have been measured in DNA of livers, kidneys and brains of wild type and knockout animals. The results revealed significant accumulation of

FapyAde in all three organs of neil1–/– and neil1–/–/nth1–/– mice, and in kidneys of nth1–/–

mice, and that of FapyGua in livers and kidneys of neil1–/– and neil1–/–/nth1–/– mice, but not in

40 brains. In contrast, no accumulation of 8-OH-Gua has been observed in these organs of any

knockout animals. Enhanced levels of FapyAde and FapyGua in knockout animals is on a par

with the substrate specificity of NEIL1 observed in vitro using DNA containing multiple

lesions (see above), and provide additional evidence for FapyAde and FapyGua, but not for

8-OH-Gua, to be the in vivo substrates of NEIL1. Significant accumulation of FapyAde and

FapyGua in cancer-prone knockout mice strongly suggests a role for these compounds in

carcinogenesis, and for the involvement of NEIL1 and NTH1 in cancer prevention. The

absence of 8-OH-Gua accumulation and GGT → GTT transversions of codon 12 in the K-ras

oncogene, which is typical of tumors in ogg1–/– and muty–/– mice, unequivocally excludes the

involvement of 8-OH-Gua in the tumor incidences observed in neil1–/– and nth1–/– animals.

There is now compelling evidence that NEIL1 is not simply a so-called backup enzyme for

other DNA glycosylases as had been assumed originally, but plays an important role in the

prevention of cancer and metabolic syndrome-associated diseases. DNA glycosylase activity

of NEIL1, which is quite distinct from those of most other known DNA glycosylases, and its

potential role in NER makes it a unique DNA repair enzyme. In addition, NEIL1 may play a

primary role in transcription- and replication-coupled repair [278,300]. Moreover, the ability of NEIL1 along with NEIL3 in the prevention of mutagenesis in vivo has recently been demonstrated [293]. The expression of NEIL1 or NEIL3 in an E. coli fpg mutY nei mutant strain significantly reduced frequency of spontaneously occurring high G → T transversions.

A greater level of FapyGua has been observed in the fpg mutY nei mutant than in the wild

type strain. The expression of NEIL1 or NEIL3 in the mutant strain significantly reduced the

level of FapyGua, confirming the specificity of NEIL1 for in vivo repair of FapyGua and

providing the evidence that NEIL3 also recognizes FapyGua in vivo. The decrease in both the

mutation frequency and the level of FapyGua suggested that the G → T transversions resulted

from FapyGua to a great extent. In fact, FapyGua has been shown to cause this type of

41 mutations [357,359]. The lack of NEIL1 activity on 8-OH-Gua in vitro and in vivo provided the evidence that 8-OH-Gua does not contribute to the adverse effects of NEIL1 deficiency in vivo. Taken together all the works surveyed above, further studies are warranted on the role of the NEIL1 substrates, FapyAde, FapyGua, R-cdA and S-cdA in carcinogenesis and

metabolic syndrome observed in neil1–/– animals.

Any role of NEIL2 or NEIL3 in carcinogenesis has not yet been investigated in detail.

Three polymorphic variants of NEIL2 have been found in patients with familial colorectal

cancer [546]. Two of these variants have also been detected in multiple colorectal carcinomas,

but also in controls and an additional variant has been found in a patient, but not in controls

[516]. A number of NEIL3 variants have been detected in multiple colorectal adenomas with

only one variant being present in a patient, but not in controls [516]. Polymorphic variants of

NEIL2 and NEIL3 have been evaluated neither for their action nor for their association with

any disease [477]. Further work will be necessary to elucidate the role of NEIL2 and NEIL3

deficiencies in carcinogenesis.

5.1.3. NTH1

Nth1–/– mice have been generated to study the effect of deficiencies in the nth1 gene

[250,251,547,548]. Several studies showed no phenotypic abnormalities in nth1–/– mice

[547,548]. Mice have been viable and exhibited similarity to wt mice in early life, and have

shown no tumor formation or increased phenotypic aberrations. No carcinogenesis has been

observed in nth1–/–/ogg1–/– mice, either [548,549]. These findings indicate that other repair

enzymes may compensate for the lack of NTH1. In humans, altered expression of the nth1

gene has recently been detected in eight gastric cancer lines [550]. Reduced mRNA

expression of NTH1 and its abnormal cytoplasmic localization have also been observed in

some primary gastric cancers. These findings pointed to a possible involvement of NTH1

deficiency in gastric cancer. Furthermore, two polymorphisms have been found in the nth1

42 promoter region; however, no association between these polymorphisms and gastric cancer has been observed. Cytoplasmic localization of NTH1 has also been detected in some primary colorectal cancers[551]. Low expression of NTH1 in the nucleus due to cytoplasmic localization in cancer cells may lead to accumulation in the nucleus of oxidatively induced

DNA lesions that are the substrates of NTH1. In another study, nth1–/– male and female mice

developed pulmonary and hepatocellular tumors as they aged; however, double knockout

animals with nth1–/–/neil1–/ exhibited a dramatic increase in both cancer incidences [251].

5.2. Other BER proteins

5.2.1. APE1

In mammalian cells, APE1, which is the mammalian ortolog of E. coli exonuclease III

family of endonucleases, provides over 95% of the total AP endonuclease function

[304,305,552-555]. In addition, APE1 exhibits multiple functions including 3'-

phosphodiestrease, 3'-5' exonuclease, 3'-phosphatase and nucleotide incision repair (NIR)

activities, transcription and redox regulations, involvement in RNA repair and metabolism

[555-559]. Critical nature of APE1 functions is evidenced by early embryonic lethality in

mice with both deleted alleles of ape1, and by increased oxidative stress, spontaneous

mutagenesis and cancer incidences, and reduced survival of pups and embryos in APE1

heterozygous mice [560-564]. Other adverse effects are caused by depletion, inhibition or

downregulation of APE1, and defects in its activity; these include apoptosis [565,566],

sensitization to DNA-damaging agents [567], and loss of neuronal function and development

of neurodegenerative disease [568-570]. There is evidence for the association of APE1

polymorphisms with disposition to cancer [571,572]. Various variants of APE1 have been

identified in the human population with the potential to lead to variations in protein activity

or expression level [485,555,573-575] (see also the NCBI database, www.ncbi.nlm.nih.gov).

43

The majority of the amino acid substitutions is located in the repair domain of the protein, whereas the redox regulatory portion (REF-1) contains several of the substitutions

[575]. The most common Asp148Glu variant has been shown to have an allele frequency of ≈

0.38 [485,573,574]. The involvement of this variant has been suggested in cancer risk such as

melanoma [576-578], [579], breast cancer [580], colorectal cancer [581-583] and

amyotropic lateral sclerosis (ALS) [584], and in ionizing radiation sensitivity [585]; however,

some other studies found no cancer susceptibility [586-590]. On the other hand, Asp148Glu

variant has been shown to possess a normal endonuclease activity [573]. Other variants

Leu104Arg, Glu126Asp and Arg237Ala exhibited a reduction in the endonuclease activity up

to ≈ 60%, while the activities of Gly241Arg and Gly306Ala variants were similar to that of

wt-APE1 [573]. Arg237Cys and Pro112Leu variants have been identified in 3 of 20

endometrial tumors [485,574]. The former displayed significant defects in several protein

activities including exonuclease function, possibly representing a reduced-function

susceptibility allele, while the latter had wt-APE1 activity, thus it seems unlikely to be involved in carcinogenesis [575]. Variants Gln51His, Ile64Val, Asp148Glu, Pro311Ser,

Gly241Arg and Ala317Val, which are not associated with human disease, exhibited no effect on the protein structure/function [575]. In terms of the intracellular localization, various

APE1 variants displayed a pattern similar to that of wt-APE1 [575]. The search in disease-

associated variations in ape1 in a large number of cancer cell lines found no novel APE1

amino acid substitutions with only common Asp148Glu and Gln51His variants observed

[575].

5.2.2. Pol β

Pol β is found in all vertebrate species and belongs to the X-family of DNA polymerases

[591-593]. In BER, following the action of APE1 that leaves one nucleotide gap with a 3'-OH and a 5'-terminal dRP, Pol β binds to the gap, performs DNA synthesis with its DNA

44 polymerase activity, filling in the gap, and removes the blocking dRP-moiety with its dRP

lyase activity, paving the way for DNA ligase 1 or a complex of X-ray repair complementing

protein 1 (XRCC1) and DNA ligase 3 to seal the resulting gap to complete the repair

[592,594-596]. The polymerase activity of Pol β is also necessary in long-patch BER

[597,598]. Pol β lacks 3' to 5' proofreading exonuclease activity and displays a moderate

fidelity with ≈ 1 error/3000 nucleotides synthesized, which is much higher than that of other

DNA polymerases [599-601]. This makes Pol β a relatively error prone polymerase.

Misinsertions introduced by Pol β during its DNA polymerase activity in BER may generate

mutations, indicating that Pol β may play a role in the etiology of cancer [602,603]. On the

other hand, Pol β appears to be essential for survival and fetal development as pol β‒/‒

animals exhibit embryonic lethality [592,604]. Mutations in pol β have been identified in

several human carcinomas and mouse lymphomas [605-611]. Variants of Pol β have been

found in approximately 30% of human tumors that also contained wt-Pol β; however, these

variants were not present in normal tissues of the same patients [602,606-608,610-615].

Approximately half of the tumors have been found to express Pol β variants with single

amino acid substitutions, whereas one third of the tumors expressed a deletion variant with

missing amino acids in the positions between 208 and 236 (Pol βΔ208-236) that likely results from alternative splicing [616,617]. This variant has been identified in several cancers

[610,612,613]; however, it has also been detected in normal tissues [612,616]. In addition, about 10% of the tumors have been shown to contain a truncated form of Pol β due to frame shift mutations. All these Pol β variants may cause defects in BER by synthesizing DNA with low fidelity, leading to genomic instability, thus to mutations leading to cancer. These mutations are not among the common polymorphisms found in pol β [574,602]. A strong association of overall survival with single nucleotide polymorphisms of pol β, i.e., the pol β

A165G and T2133G phenotypes, has been found in a large number of patients with

45 pancreatic adenocarcinoma [618]. In contrast, the homozygous variant genotype of pol β had

a significant protective effect on overall survival. The median survival time was 35.7 months

for patients with at least one of the two homozygous variant pol β 165 GG and 2133 CC

genotypes, whereas those patients carrying the pol β 165 AA/AG and 2133 TT/TC genotypes

had a median survival time of 14.8 months.

A Pol β Lys289Met variant along with wt-Pol β has been identified in a human colon

carcinoma [612]. The expression of this variant in mouse cells led to a significant increase in

the mutation frequency with the mutational spectrum being different from that of wt-Pol β

[619]. In addition, the Lys289Met variant displayed a lower fidelity than wt-Pol β and

misincorporated nucleotides during BER. The frequency of C → G transversion mutations in

cells with Lys289Met variant expression was much greater than those in wild type cells.

Another Ile260Met variant of Pol β has been identified in a prostate carcinoma [606]. This

variant has been shown to have a mutational spectrum different from that of wt-Pol β and to possess a sequence-specific mutator activity [620,621]. Expression of the Lys289Met and

Ile260Met variants in mouse cells has been shown to result in permanent cellular

transformation, indicating that these variants induce mutations during BER by aberrant gap filling, which is likely to be different from the function of wt-Pol β [620]. These results led to the suggestion that the mutations may occur in key growth control cells leading to cellular transformation. By genotyping of pol β, two exonic germline variants Arg137Gln and

Pro242Arg of Pol β have been found with allele frequencies of 9% and 3%, respectively

[622]. The Arg137Gln variant displayed a lower DNA polymerase activity than wt-Pol β

[623]. Patients with lung cancer who carry the Pro242Arg variant exhibited a decrease in survival [624]. A gastric cancer-associated variant Glu295Lys of Pol β has been shown to have no DNA polymerase activity, leading to unfilled gap in BER and inducing cellular transformation [625]. Another gastric cancer-associated Leu22Pro variant exhibited DNA

46 polymerase activity, albeit lesser than that of wt-Pol β, but lacked dRP lyase activity and

could not support BER [626]. The studies of Pol β mutants lent credence to the mutator

phenotype hypothesis [15] and suggested that BER is a tumor suppressor mechanism [603].

Pol β has been shown to form complexes with DNA lesions such as mutagenic 8-OH-Gua

in the confines of its active site [592,627]. It preferentially inserted the correct base cytosine

opposite 8-OH-Gua rather than adenine, although this depended on the sequence context and

the insertion of adenine also occurred [628,629]. However, 8-OH-Gua has been the only

DNA lesion investigated so far. Pol β may also form complexes with other oxidatively

induced DNA base lesions, causing misinsertions and subsequent mutations. Taken together,

accumulated evidence clearly points to a role of Pol β and its variants in the etiology of

cancer.

5.3. DNA lesions and DNA repair proteins as biomarkers

5.3.1.DNA lesions as biomarkers

Evidence accumulated over several decades suggests that oxidatively induced DNA

lesions may be used as potential disease biomarkers and cancer risk assessment. Elevated

levels of such DNA lesions in cancerous tissues and in BER enzyme-knockout animals or in

animals that developed cancer upon exposure to environmental toxins, as discussed above,

strongly support this notion. For this purpose, however, accurate measurements of oxidatively

induced DNA lesions in tissues would be absolutely necessary. Mass spectrometry-based

assays using stable isotope-labeled internal standards have been developed for such

measurements and successfully applied over the past two decades or so. It is out of the scope

of this article to review the entire literature on this subject. The reader is referred to reviews

and other articles in this field (see e.g., [105,121,124,251,630]). Noninvasive procedures to

collect samples such as urine and the measurement of oxidatively induced DNA lesions

47 therein drew significant attention from many laboratories. A large number of studies have been conducted to measure these lesions as non-invasive biomarkers for diagnosis, early detection and therapy monitoring, and also for epidemiological studies. Various protocols have been developed to measure DNA lesions in urine. These included the use of mass spectrometry, HPLC with electrochemical detection and enzyme-linked immunosorbant assay

(ELISA). First studies measured thymine glycol and 2'-deoxythymidine glycol (dT glycol)

[631], and 8-hydroxy-2'-deoxyguanosine (8-OH-dG) [632]. Subsequently, 8-OH-dG and its free base 8-OH-Gua, albeit to a lesser extent, have mainly been measured [633-656]. Other lesions FapyGua, 8-OH-Ade and 5-OH-Ura have also been found in urine [633,644], although much less attention was paid to these lesions than 8-OH-dG and 8-OH-Gua. There have been significant differences between the measurements in different laboratories and about the source of these DNA lesions in urine. A European Standards Committee on Urinary

(DNA) Analysis (ESCULA) has been established to help reach a consensus between results

in different laboratories in Europa, USA and Asia [649]. A recent comprehensive study

involving a large number of laboratories investigated both human and methodological factors

influencing measurements of 8-OH-dG in urine using ELISA and chromatographic

techniques including mass spectrometry [656]. Chromatographic techniques performed better

than ELISA in terms of high agreement across urine samples from different subjects. As for

the source, several studies showed that diet and cell death do not contribute to the appearance

of DNA lesions in urine [635,639,645,647,649,657,658]. BER has been suggested to be

responsible for the presence of 8-OH-Gua in urine because of the specificity of OGG1 of

BER for this lesion [640]. However, DNA glycosylases such as OGG1 would not remove 8-

OH-dG from DNA because they are specific for modified DNA bases, not for modified 2'-

deoxynucleosides [217]. NER is not likely to be responsible for 8-OH-dG in urine, either,

because no oligonucleotides containing 8-OH-dG have been identified in urine [639]. Repair

48 in the nucleotide pool appears to be a major source of 8-OH-dG and possibly that of dT glycol and other modified 2'-deoxynucleosides [657,659]. MTH1 is the best characterized enzyme in the nucleotide pool that hydrolyzes 8-OH-dGTP to 8-OH-dGMP and thus prevents its incorporation into DNA [197,660]. Dephosphorylation of the latter would give rise to 8-

OH-dG, which might be removed from cells and ultimately appear in urine. Still, the precise nature of the presence of 8-OH-dG in urine remains unclear [661].

Recently, the presence of R-cdA and S-cdA has been discovered in human urine [127]. A methodology has been developed to simultaneously measure these compounds and 8-OH-dG

in urine using LC-MS/MS with isotope-dilution. Since 8,5'-cyclopurine-2'-deoxynucleosides are repaired by NER, not by BER [108,294-296], their presence in urine has been suggested to result from this repair pathway[127]. However, the repair of their triphosphates in the nucleotide pool followed by the excretion into urine as described above cannot be excluded, either. In an application to a disease state, R-cdA and S-cdA have been found in urine of atherosclerosis patients at significantly greater concentrations than in that of healthy individuals [128] (Fig. 8). The statistical difference was highly significant. 8-OH-dG has been simultaneously measured. Its concentrations in urine of patients were also significantly greater than those in controls. However, the significance of the data for R-cdA and S-cdA was greater than that for 8-OH-dG. The concentrations of R-cdA and S-cdA in urine were about two magnitudes of order less than that of 8-OH-dG. 8-OH-Gua has also been measured by another method. No statistical significance has been found between the levels of 8-OH-Gua in controls and patients, indicating that this compound may not be a reliable biomarker.

Simultaneous measurement of R-cdA and S-cdA along with 8-OH-dG in urine would be more advantageous for reliable results than the measurement of one lesion only. Taken together, the accurate and reproducible measurement of R-cdA and S-cdA in human urine, and their extraordinary chemical stability and clear origin render these compounds ideal as potential

49 disease biomarkers in urine for early detection, testing of drugs, monitoring the therapy and epidemiological studies. Further studies on the use of these unique compounds as biomarkers for cancer and other diseases are warranted.

5.3.2. BER proteins as biomarkers

Ionizing radiation and most chemotherapeutic agents kill tumor cells by damaging DNA.

However, their effectiveness may be influenced by the efficiency of DNA repair capacity in

tumors [449,454,480,662-664]. Overexpression of DNA repair proteins that may increase the

DNA repair capacity is common in cancer. Since malignant tumors possess increased level of

oxidatively induced DNA damage [112,113,116,117,457-470], the overexpression of DNA

repair proteins may be required to encounter high level of DNA damage in tumors. Increased

rate of DNA damage and subsequent mutations may lead to genetic instability and cell death

late in tumor evolution. However, rapidly developing tumors that overexpress DNA repair

proteins may have an evolutionary advantage for survival and thus develop greater DNA

repair capacity than normal tissues. Effective DNA repair in tumors that removes DNA

lesions before they become toxic is a major mechanism for resistance to therapy, and may

affect the outcome of therapy and thus determine the patient survival. In the normal

population, cancer susceptibility may also be influenced by repair capacity [449]. Increase in

DNA repair capacity allows cancer cells to develop multi-drug resistance. Accumulated

evidence strongly suggests that DNA repair capacity might be a predictive biomarker of

patient response to therapy [480,663]. Determination of the overexpression or

underexpression of DNA repair proteins in normal and cancer tissues might help predict and

guide treatments. In this context, BER proteins have emerged as biomarkers for prediction of

tumor response and prognosis of treatment outcome [449,454,479,480,555,572,663,665-667].

50

5.3.3. BER proteins as therapy targets

Since BER proteins are responsible for the repair of a multiplicity of oxidatively induced

DNA lesions, they would be logical targets for inhibition to effectively achieve therapies to

overcome the resistance of tumors to treatment, effecting apoptosis or cell death instead of

DNA repair [555]. Development of inhibitors of BER proteins and other DNA repair proteins

for combination therapy or as single agents for monotherapy will help selectively kill tumors.

The achievement of this goal will potentially lead to personalized cancer therapy [480]. In

this respect, a thorough understanding of DNA repair pathways and functions of proteins

involved will be of fundamental importance for the use of DNA repair proteins as biomarkers

and targets for improving therapy of cancer [668].

Success with the inhibitors of poly(ADP ribose) polymerase 1 (PARP1), which is the major enzyme of the members of the PARP superfamily and involved in BER, brought to attention the inhibition of BER proteins in tumors as a potential and promising concept for cancer therapy (reviewed in [667]). Inhibition of PARP activity preventing the repair of methylation-induced DNA strand breaks and generating cytotoxicity gave the impetus for development of DNA repair inhibitors for cancer therapy [669]. In particular, high efficacy of

PARP inhibitors as single agents for monotherapy has been demonstrated in patients with

inherited breast and ovarian cancers that contain deficient and brca2 genes [670,671].

About a decade ago, PARP inhibitors started entering clinical trials [672,673]. Clinical trials

are in progress now and numerous PARP inhibitors that have been developed by the

academia and the industry are being tested. There is a wealth of literature dealing with PARP

inhibitors and their use in cancer therapy. More information can be found in several recent

review articles [454,663,667,674,675].

APE1 is another BER protein for which intense efforts are being underway worldwide to

find inhibitors of its activities [555]. Overexpression of APE1 has been observed in multiple

51 human cancers and associated with resistance to and radiation therapy, and with poor survival[555,571,572,676]. In this context, numerous studies provided the evidence for the expression and also subcellular localization of APE1 to be of great predictive and prognostic value. Thus, strict nuclear localization of APE1 has been found to associate with good prognosis, whereas its combined nuclear and cytoplasmic localization correlated with poor survival. However, elevated levels of APE1 with predominantly nuclear localization have also been found in various cancers [677-681]. These contradictory observations indicate that APE1 subcellular localization may vary among cancer types. Mounting evidence suggests that inhibition of APE1 functions increases cellular sensitivity to DNA-damaging agents. As discussed above, this protein is not only a DNA repair enzyme, but also exhibits other important functions and aspects. Because of these reasons, APE1 is generally accepted as being an excellent target for development of inhibitors as anticancer agents for monotherapy and/or to enhance the efficacy of present drugs and ionizing radiation in cancer therapy [555,572,682]. Since 2005, thousands of structurally diverse compounds have been screened to find inhibitors of APE1 using high-throughput screens [471,555,572,682-686].

Inhibitors that block the DNA repair function of APE1 work either by binding to DNA to

inhibit its endonuclease activity or by binding to APE1 to inhibit its AP site activity. Some

inhibitors of APE1 endonuclease function are in Phase I trials [555]. Inhibition of the redox

activity of APE1 blocks various cellular pathways including multiple tumor signaling

pathways involved in cancer development and survival [687,688]. There are naturally-

occurring or synthesized inhibitors of the redox activity of APE1 that are being tested or are

in development [555].

Inhibiting Pol β may also be of importance for cancer therapy. Thus, greater levels of Pol

β have been found in breast, colon and prostate adenocarcinomas than in adjacent normal

tissues[689]. Overexpression of Pol β, which decreases the fidelity of BER[690], has been

52 associated with genetic instability, cellular transformation, hyperplasia and carcinogenesis as

well as resistance to therapy with DNA-damaging agents [690-695]. In fact, Pol β has been

shown to be upregulated in the presence of increased DNA damage [696,697]. Knocking

down Pol β has increased cellular sensitivity to DNA-damaging agents [698,699]. Efforts are

being underway to develop small molecule inhibitors of Pol β activities. One of the first

inhibitors exhibited inhibition of both lyse and polymerase activities of Pol β [700].

Numerous other inhibitors have been developed with a variety of potencies and specificities

[700-703]. Many of them lacked necessary characteristics to become cancer-specific drugs.

However, some of these compounds were promising and are in preclinical studies [555]. One

compound enhanced the ability of DNA-alkylating agent (TMZ), which has

been used successfully for the treatment of some cancers [704,705], to impair the growth of

colon cancer cells [702]. Another study tested thousands of small molecules targeting Pol β

for chemotherapeutic intervention of colorectal cancer [706]. A compound with a low

molecular weight has been identified to be a potential inhibitor of Pol β activities. However,

it did not affect the activity of other BER proteins. Combining this small molecule inhibitor

with TMZ effectively blocked the growth of colon cancer cells in vitro and caused antitumor

activity in vivo. Taken together, more efforts may be needed to study mechanistic aspects of

Pol β activities and to develop high-throughput screening assays for the search of inhibitors

of this important BER protein to be used as anti-cancer drugs.

Flap endonuclease I (FEN1) is another BER protein, for which there are efforts to develop

inhibitors. This protein is involved in long-patch BER [707-712]. Following the action of

APE1, Pol δ/ε introduces two to eight deoxynucleotides past the AP site generating an

overhang polydeoxynucleotide with two to ten deoxynucleotides (5' flap) and the dRP. FEN1

removes this 5' flap and then DNA ligase I seals the remaining nick completing the repair.

There are other functions of FEN1 as well [711-714]. Efficient activities of FEN1 are

53 essential for the maintenance of the genomic integrity [715]. Overexpression of FEN1 has

been observed in numerous cancers, suggesting a role for FEN1 in tumor progression and

development, and therapy resistance [716-722]. Consequently, FEN1 might be a potential target for anti-cancer treatment. Assays have been developed to find inhibitors of FEN1

[723,724]. Several small molecule inhibitors with N-hydroxyurea-based compounds among them have been identified with the potential to serve as anti-cancer drugs [723-725]. FEN1 has also been found to be a potential target for synthetic lethality [726]. Some N-hydroxyurea series of FEN1 inhibitors are in early preclinical trials [555]. Developed assays may facilitate the developments of novel inhibitors of FEN1 to be used in cancer therapy.

Recent work suggested that targeting MTH1 of the nucleotide pool repair protein might

be beneficial in cancer therapy [204,727]. This protein dephosporylates modified 2'-

deoxynucleoside triphosphates in the nucleotide pool, preventing their incorporation into

DNA during DNA replication [197-202]. Modified 2'-deoxynucleoside triphosphates in the

nucleotide pool have been shown to be a significant contributor to genetic instability in

mismatch repair-deficient cells [728]. Overexpression of MTH1 has been observed in many cancers [200,203,729,730]. MTH1 activity has been found to be greater in tumors than surrounding normal tissues from non-small-cell lung cancer patients [203]. The level of 8-

OH-Gua was also lower in tumors than in surrounding normal tissues, suggesting DNA repair capacity of tumors may be greater that that of normal tissues. This is in agreement with reported lower levels of oxidatively induced DNA lesions and ethano-DNA adducts in cancerous tissues compared to surrounding non-cancerous tissues [123,481,482]. Cancer cells have been shown to require MTH1 for efficient survival by avoiding incorporation of modified 2'-deoxynucleoside triphosphates into their DNA during replication [204]. This finding suggested that MTH1 might be an excellent target for inhibition in cancer treatment.

Small molecule inhibitors of MTH1 have been found by screening compound libraries and

54 successfully applied to suppress tumor growth in mice with different cancers, validating

MTH1 as a novel anticancer target in vivo. Clinical trials of MTH1 inhibitors may be

performed in the future for drug development.

Despite the successes with other BER proteins, the development of inhibitors for DNA

glycosylases has been lagging. Recently, a study has been conducted to identify gene-specific

pathways that would serve as synthetic lethal partners with DNA glycosylases as the targets

for cancer therapy using chemotherapeutic agents that function through depletion of cellular

nucleotide pools [731]. Thymidylate synthetase (TS) plays the key role in the synthesis of 2'-

deoxythymidine (dT) by producing dTMP from dUMP [732,733]. Inhibitors that target the

TP pathway are widely used in the treatment of many cancers [734-737]. These are mainly

folate-based analogues with some nucleotide-based compounds and cause toxicity by

depletion of dTTP that inhibits DNA replication and increases incorporation of dUTP into

DNA. The combined siRNA-mediated reduction of NEIL1 and the treatment with four TS

inhibitors dramatically increased toxicity in an osteosarcoma cell line [731]. Depletion of

NEIL1 or OGG1 alone had no effect on toxicity. Moreover, loss of NEIL function has been shown to be synthetically lethal with the disruption of the DNA repair pathway. This work identified NEIL1 as the key BER protein in the repair of DNA following inhibition of TS pathway. This means that inhibition of NEIL1 may enhance clinical responses to TS pathway inhibitors. A recent work developed a strategy to discover inhibitors of DNA glycosylases with the goal of finding small molecules to be used in combination

therapy [738]. A high-throughput, fluorescence-based assay has been developed that uses incision of oligodeoxynucleotides with a single modified DNA base to detect small molecule inhibitors of DNA glycosylases with an associated AP lyase activity. NEIL1 has been used as the proof-of-principle glycosylase for this purpose. Oligodeoxynucleotides contained both Sp and Gh that are oxidation products of 8-OH-Gua (see above). As discussed above, NEIL1 is

55 specific for excision of FapyAde and FapyGua from DNA in vitro and in vivo, and also

efficiently excises Sp and Gh from oligodeoxynucleotides, but it does not act on 8-OH-Gua.

The high-throughput assay has been used to screen small molecule libraries with a large

number of compounds for inhibitors of the combined glycosylase/AP lyase activities. There

were a number of purine analogues among top hits of these screens. Since FapyAde and

FapyGua are physiological substrates of NEIL1, inhibition by purine analogues of NEIL1

activity for these compounds has been tested using DNA samples that contained multiple

oxidatively induced lesions. Fig. 9 illustrates the determination of the inhibition of NEIL1

activity for FapyAde and FapyGua by six purine analogues. An efficient activity of NEIL1 on

FapyAde and FapyGua has been observed in agreement with previous studies

[250,251,278,290,292]. Inhibitors P2, P6, P7 and P8 significantly (p < 0.005) inhibited the

excision of the two lesions, whereas P11 had a lesser but significant effect on FapyAde

excision and did not inhibit FapyGua excision. Taken together, this study may form a foundation for potential drug discovery for cancer therapy for the entire family of DNA glycosylases that are specific for a variety of oxidatively induced DNA base lesions. As was done in the study of NEIL1 inhibitors, simultaneous measurement of biological substrates of these enzymes using DNA samples with multiple DNA lesions will facilitate screening of

various potential inhibitors with the goal of discovering suitable drugs for inhibition of DNA

glycosylases in cancer therapy.

As Kelley has stated [663], knowledge of repair proteins’ overexpression or

underexpression in cancers may help predict and guide development of treatments,

potentially yielding the greatest therapeutic response. Thus far, the expression of DNA repair proteins in cells and tissues including clinical samples have been estimated by semi-

quantitative immunochemical methods. To be used as reliable biomarkers in cancer,

expression levels of DNA proteins must be accurately measured in tissues by proper chemical

56 and physical techniques. Mass spectrometry is the most suitable technique of choice for this

purpose and is being used worldwide for the measurement of proteins in the field of

proteomics. The application of this technique would be essential for positive identification

and accurate quantification of DNA repair proteins in human tissues. Recently, our laboratory

has developed methodologies that use mass spectrometry with isotope-dilution for the

measurements of DNA repair proteins [739-742]. Full length 15N-labeled analogues of human

OGG1, NEIL1, NTH1, Pol β and APE1 have been produced and purified to be used as internal standards for the accurate quantification of these proteins by LC-MS/MS. Thus,

APE1 has been identified and quantified in human tissues, cultured human cells and mouse liver. Efforts are now being made to extend this work to measure expression levels of DNA proteins in different types of human cancer tissues from patients and in tissues from disease- free individuals. We believe that such measurements will be of fundamental importance for the determination of DNA repair capacity, the use of DNA repair proteins as biomarkers and

the development of DNA repair inhibitors.

6. Conclusions

There is mounting evidence for an important role of oxidatively induced DNA damage

and its cellular repair in the etiology of cancer. Great strides have been made in the

understanding of various mechanisms of DNA repair since its discovery five decades ago.

Discovery of DNA repair proteins and elucidation of their functions paved the way for

understanding of the contribution of aberrant DNA repair, and mutations and polymorphisms

in DNA repair genes to carcinogenesis. The finding that increased DNA repair capacity in tumors causes resistance to therapy gave impetus to development of inhibitors of DNA repair to increase the efficacy of therapy. Success of the application of certain inhibitors as drugs in the therapy of some cancers is a very promising development. It will be important to continue

57 the development of inhibitors of proteins in numerous DNA repair pathways as drugs to be used to enhance the efficacy of current cancer therapy and to eradicate cancer. A thorough understanding of the DNA repair pathways in carcinogenesis and overexpression or

underexpression of DNA repair proteins in tumors in comparison to normal tissues will be of

utmost importance in improving therapy and achieving the best therapeutic response.

Conflict of interest statement

The author declares that there are no conflicts of interest.

58

References

[1] B. Halliwell, J.M.C. Gutteridge. Free Radicals in Biology and Medicine, Fourth Edition, Oxford University Press, Oxford, 2007. [2] H. Sies. Oxidative Stress: Oxidants and Antioxidants, Academic Press, London, 1991. [3] D. Hanahan, R.A. Weinberg. The hallmarks of cancer, Cell 100 (2000) 57-70. [4] L.M. Coussens, Z. Werb. Inflammation and cancer, Nature 420 (2002) 860-867. [5] A. Mantovani, P. Allavena, A. Sica, F. Balkwill. Cancer-related inflammation, Nature 454 (2008) 436-444. [6] M. Dizdaroglu, R. Olinski, J.H. Doroshow, S.A. Akman. Modification of DNA bases in chromatin of intact target human cells by activated human polymorphonuclear leukocytes, Cancer Res. 53 (1993) 1269-1272. [7] A.B. Ross, W.G. Mallard, W.P. Helman, B.H.J. Bielski, G.V. Buxton, D.E. Cabelli, C.L. Greenstock, R.E. Huie, P. Neta. NDRL-NIST solution kinetics database, National Institute of Standards and Technology, Gaithersburg, 1992. [8] R.E. Huie, S. Padmaja. The reaction of no with superoxide, Free Radic. Res. Commun. 18 (1993) 195-199. [9] C. von Sonntag. Free-Radical-Induced DNA Damage and Its Repair, Springer, Hiedelberg, 2006. [10] M. Dizdaroglu, P. Jaruga. Mechanisms of free radical-induced damage to DNA, Free Radic. Res. 46 (2012) 382-419. [11] E.C. Friedberg, G.C. Walker, W. Siede, R.D. Wood, R.A. Schultz, T. Ellenberger. DNA Repair and Mutagenesis, ASM Press, Washington, D.C., 2006. [12] B. Vogelstein, K.W. Kinzler. The Genetic Basis of Human Cancer, McGraw-Hill, New York, 1998. [13] E.C. Friedberg. DNA damage and repair, Nature 421 (2003) 436-440. [14] R.A. Beckman, L.A. Loeb. Genetic instability in cancer: theory and experiment, Semin. Cancer Biol. 15 (2005) 423-435. [15] L.A. Loeb. Human cancers express mutator phenotypes: origin, consequences and targeting, Nature Rev. Cancer 11 (2011) 450-457. [16] C.H. Bosken, Q. Wei, C.I. Amos, M.R. Spitz. An analysis of DNA repair as a determinant of survival in patients with non-small-cell lung cancer, J. Natl. Cancer Inst. 94 (2002) 1091-1099. [17] R.V. Lord, J. Brabender, D. Gandara, V. Alberola, C. Camps, M. Domine, F. Cardenal, J.M. Sanchez, P.H. Gumerlock, M. Taron, J.J. Sanchez, K.D. Danenberg, P.V. Danenberg, R. Rosell. Low ERCC1 expression correlates with prolonged survival after plus chemotherapy in non-small cell lung cancer, Clin. Cancer Res. 8 (2002) 2286-2291. [18] R. Rosell, M. Taron, A. Barnadas, G. Scagliotti, C. Sarries, B. Roig. Nucleotide excision repair pathways involved in Cisplatin resistance in non-small-cell lung cancer, Cancer Control 10 (2003) 297-305. [19] L.-E. Wang, Q. Wei. Nucleotide excision repair: DNA repair capacity, variability and cancer susceptibility, in: S. Madhusudan and D.M. Wilson, III (Eds.), DNA Repair and Cancer From Bench to Clinic, CRC Press, Boca Raton, 2013, pp. 288- 309. [20] A.L. Jackson, L.A. Loeb. The contribution of endogenous sources of DNA damage to the multiple mutations in cancer, Mutat. Res. 477 (2001) 7-21. [21] J.F. Davidson, H.H. Guo, L.A. Loeb. Endogenous mutagenesis and cancer, Mutat. Res. 509 (2002) 17-21.

59

[22] M. Dizdaroglu. Oxidatively induced DNA damage: Mechanisms, repair and disease, Cancer Lett. 327 (2012) 26-47. [23] S. Steenken. Purine bases, nucleosides, and nucleotides: Aqueous solution redox chemistry and transformation reactions of their radical cations and e- and OH adducts, Chem. Rev. 89 (1989) 503-520. [24] A. Hissung, S.C. von. The reaction of solvated electrons with cytosine, 5-methyl cytosine and 2'-deoxycytidine in squeous solution. The reaction of the electron adduct intermediates with water, p-nitroacetophenone and oxygen. A pulse spectroscopic and pulse conductometric study, Int. J. Radiat. Biol. Relat Stud. Phys. Chem. Med. 35 (1979) 449-458. [25] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms, and hydroxyl radicals (•OH/•O_) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988) 513-886. [26] L.P. Candeias, P. Wolf, P. O'Neill, S. Steenken. Reaction of hydrated electrons with guanine nucleosides: Fast protonation on carbon of the electron adduct, J. Phys. Chem. 96 (1992) 10302-10307. [27] M. D'Angelantonio, M. Russo, P. Kaloudis, Q.G. Mulazzani, P. Wardman, M. Guerra, C. Chatgilialoglu. Reaction of hydrated electrons with guanine derivatives: tautomerism of intermediate species, J. Phys. Chem. B 113 (2009) 2170-2176. [28] P. O'Neill. Pulse radiolytic study of the interaction of thiols and ascorbate with OH adducts of dGMP and dG: Implications for DNA repair processes., Radiat. Res. 96 (1983) 198-210. [29] C. Chatgilialoglu, M. D'Angelantonio, M. Guerra, P. Kaloudis, Q.G. Mulazzani. A reevaluation of the ambident reactivity of the guanine moiety towards hydroxyl radicals, Angew. Chem. Int. Ed Engl. 48 (2009) 2214-2217. [30] C. Chatgilialoglu, M. D'Angelantonio, G. Kciuk, K. Bobrowski. New insights into the reaction paths of hydroxyl radicals with 2'-deoxyguanosine, Chem. Res. Toxicol. 24 (2011) 2200-2206. [31] C.J. Mundy, M.E. Colvin, A.A. Quong. Irradiated guanine: A Car-Parrinello molecular dynamics study of dehydrogenation in the presence of an OH radical, J. Phys. Chem. 106 (2002) 10063-10071. [32] Y. Wu, C.J. Mundy, M.E. Colvin, R. Car. On the mechanisms of OH radical induced DNA-base damage: A comparative quantum chemical and Car-Parrinello molecular dynamics study., J. Phys. Chem. 108 (2004) 2922-2929. [33] A. Adhikary, A. Kumar, D. Becker, M.D. Sevilla. The guanine cation radical: investigation of deprotonation states by ESR and DFT, J. Phys. Chem. B 110 (2006) 24171-24180. [34] S. Naumov, S.C. von. Guanine-derived radicals: dielectric constant-dependent stability and UV/Vis spectral properties: a DFT study, Radiat. Res. 169 (2008) 364- 372. [35] S.D. Phadatare, K.K. Sharma, B.S. Rao, S. Naumov, G.K. Sharma. Spectral characterization of guanine C4-OH adduct: a radiation and quantum chemical study, J. Phys. Chem. B 115 (2011) 13650-13658. [36] T.P. Troy, M. Nakajima, N. Chalyavi, K. Nauta, S.H. Kable, T.W. Schmidt. Hydroxyl Addition to Aromatic Alkenes: Resonance-Stabilized Radical Intermediates, J. Phys. Chem. A 116 (2012) 7906-7915. [37] A.J.S.C. Vieira, S. Steenken. Pattern of OH radical reactions with N6,N6- dimethyladenosine. Production of three isomeric OH adducts and their dehydration and ring-opening reactions, J. Amer. Chem. Soc. 109 (1987) 7441-7448.

60

[38] A.J.S.C. Vieira, S. Steenken. Pattern of OH radical reaction with adenine and its nucleosides and nucleotides. Characterization of two types of isomeric OH adduct and their unimolecular transformation reactions., J. Am. Chem. Soc. 112 (1990) 6986-6994. [39] L.P. Candeias, S. Steenken. Reaction of HO• with guanine derivatives in aqueous solution: Formation of two different redox-active OH-adduct radicals and their unimolecular transformation reactions. Properties of G(-H)•, Chemistry European Journal 6 (2000) 475-484. [40] M.C.R. Symons. Apllication of electron spin resonance spectroscopy to the study of the effects of ionising radiation on DNA and DNA complexes., J. Chem. Soc. , Faraday Trans. 83 (1987) 1-11. [41] S. Boiteux, E. Gajewski, J. Laval, M. Dizdaroglu. Substrate specificity of the Escherichia coli Fpg protein (formamidopyrimidine-DNA glycosylase): excision of purine lesions in DNA produced by ionizing radiation or photosensitization, Biochemistry 31 (1992) 106-110. [42] P.W. Doetsch, T.H. Zastawny, A.M. Martin, M. Dizdaroglu. Monomeric base damage products from adenine, guanine, and thymine induced by exposure of DNA to radiation, Biochemistry 34 (1995) 737-742. [43] T. La Vere, D. Becker, M.D. Sevilla. Yields of .OH in gamma-irradiated DNA as a function of DNA hydration: Hole transfer in competition with .OH formation, Radiat. Res. 145 (1996) 673-680. [44] S.G. Swarts, D. Becker, M. Sevilla, K.T. Wheeler. Radiation-induced DNA damage as a function of hydration. II. Base damage from electron-loss centers, Radiat. Res. 145 (1996) 304-314. [45] H. Kasai, S. Nishimura. Hydroxylation of the C-8 position of deoxyguanosine by reducing agents in the presence of oxygen, Nucleic Acids Symp. Ser. 12 (1983) 165-167. [46] H. Kasai, S. Nishimura. Hydroxylation of deoxyguanosine at the C-8 position by polyphenols and aminophenols in the presence of hydrogen peroxide and ferric ion, Gann 75 (1984) 565-566. [47] H. Kasai, H. Tanooka, S. Nishimura. Formation of 8-hydroxyguanine residues in DNA by X-irradiation, Gann 75 (1984) 1037-1039. [48] H. Kasai, S. Nishimura. Hydroxylation of deoxyguanosine at the C-8 position by ascorbic acid and other reducing agents, Nucleic Acids Res. 12 (1984) 2137-2145. [49] M. Dizdaroglu. Application of capillary gas chromatography-mass spectrometry to chemical characterization of radiation-induced base damage of DNA; implications for assessing DNA repair processes, Anal. Biochem. 144 (1985) 593-603. [50] M. Dizdaroglu. Formation of an 8-hydroxyguanine moiety in deoxyribonucleic acid on gamma-irradiation in aqueous solution, Biochemistry 24 (1985) 4476-4481. [51] J. Reynisson, S. Steenken. DFT calculations on the electrophilic reaction with water of the guanine and adenine radical cations. A model for the situation in DNA, Phys. Chem. Chem. Phys. 4 (2002) 527-532. [52] M. Aida, S. Nishimura. An ab initio molecular orbital study on the characteristics of 8-hydroxyguanine, Mutat. Res. 192 (1987) 83-89. [53] S.J. Culp, B.P. Cho, F.F. Kadlubar, F.E. Evans. Structural and conformational analyses of 8-hydroxy-2'-deoxyguanosine, Chem. Res. Toxicol. 2 (1989) 416-422. [54] G. Hems. Effect of ionizing radiation on aqueous solutions of guanylic acid and guanosine, Nature 181 (1958) 1721-1722. [55] G. Hems. Chemical effects of ionizing radiation on deoxyribonucleic acid in dilute aqueous solution, Nature 186 (1960) 710-712.

61

[56] M. Dizdaroglu, G. Kirkali, P. Jaruga. Formamidopyrimidines in DNA: Mechanisms of formation, repair, and biological effects, Free Radic. Biol. Med. 45 (2008) 1610- 1621. [57] Z. Nackerdien, K.S. Kasprzak, G. Rao, B. Halliwell, M. Dizdaroglu. Nickel(II)- and cobalt(II)-dependent damage by hydrogen peroxide to the DNA bases in isolated chromatin, Cancer Res. 51 (1991) 5837-5842. [58] K.S. Kasprzak, B.A. Diwan, J.M. Rice, M. Misra, C.W. Riggs, R. Olinski, M. Dizdaroglu. Nickel(II)-mediated oxidative DNA base damage in renal and hepatic chromatin of pregnant rats and their fetuses. Possible relevance to carcinogenesis, Chem. Res. Toxicol. 5 (1992) 809-815. [59] J. Cadet, M. Berger, G.W. Buchko, P.C. Joshi, S. Raoul, J.-L. Ravanat. 2,2- Diamino-4-[(3,5-di-O-acetyl-2-deoxy-beta-D-erythrosepentofuranosyl) amino]-5- (2H)-oxazolone - A novel and predominant radical oxidation product of 3',5'-di-O- acetyl-2'-deoxyguanosine., J. Am. Chem. Soc. 116 (1994) 7403-7404. [60] S. Raoul, M. Berger, G.W. Buchko, P.C. Joshi, B. Morin, M. Weinfeld, J. Cadet. H- 1, C-13 and N-15 nuclear magnetic resonance analysis and chemical features of the two main radical oxidation products of 2'-deoxyguanosine: Oxazolone and imidazolone nucleosides., J. Chem. Soc. , Perkin Trans. 2 3 (1996) 371-381. [61] C. von Sonntag. Topics in free radical-mediated DNA damage: purines and damage amplification-superoxic reactions-, the incomplete radiomimetic, Int. J. Radiat. Biol. 66 (1994) 485-490. [62] R. Misiaszek, C. Crean, A. Joffe, N.E. Geacintov, V. Shafirovich. Oxidative DNA damage associated with combination of guanine and superoxide radicals and repair mechanisms via radical trapping, J. Biol. Chem. 279 (2004) 32106-32115. [63] B. Matter, D. Malejka-Giganti, A.S. Csallany, N. Tretyakova. Quantitative analysis of the oxidative DNA lesion, 2,2-diamino-4-(2-deoxy-beta-D-erythro- pentofuranosyl)amino]-5(2H)-oxazolon e (oxazolone), in vitro and in vivo by isotope dilution-capillary HPLC-ESI-MS/MS, Nucleic Acids Res. 34 (2006) 5449- 5460. [64] J. Cadet, T. Douki, J.L. Ravanat. Oxidatively generated damage to the guanine moiety of DNA: mechanistic aspects and formation in cells, Acc. Chem. Res. 41 (2008) 1075-1083. [65] S. Steenken, S.V. Jovanovic, M. Bietti, K. Bernhard. The trap depth (in DNA) of 8- oxo-7,8-dihydro-2'-deoxyguanosine as derived from electron-transfer equilibria in aqeous solution., J. Am. Chem. Soc. 122 (2000) 2373-2384. [66] W. Luo, J.G. Muller, E.M. Rachlin, C.J. Burrows. Characterization of spiroiminodihydantoin as a product of one-electron oxidation of 8-Oxo-7,8- dihydroguanosine, Org. Lett. 2 (2000) 613-616. [67] W. Luo, J.G. Muller, E.M. Rachlin, C.J. Burrows. Characterization of hydantoin products from one-electron oxidation of 8-oxo-7,8-dihydroguanosine in a nucleoside model, Chem. Res. Toxicol. 14 (2001) 927-938. [68] J.C. Niles, J.S. Wishnok, S.R. Tannenbaum. Spiroiminodihydantoin is the major product of the 8-oxo-7,8-dihydroguanosine reaction with peroxynitrite in the presence of thiols and guanosine photooxidation by methylene blue, Org. Lett. 3 (2001) 963-966. [69] C.J. Burrows, J.G. Muller, O. Kornyushyna, W. Luo, V. Duarte, M.D. Leipold, S.S. David. Structure and potential mutagenicity of new hydantoin products from guanosine and 8-oxo-7,8-dihydroguanine oxidation by transition metals, Environ. Health Perspect. 110 Suppl 5 (2002) 713-717.

62

[70] W. Adam, M.A. Arnold, M. Grune, W.M. Nau, U. Pischel, C.R. Saha-Moller. Spiroiminodihydantoin is a major product in the photooxidation of 2'- deoxyguanosine by the triplet states and oxyl radicals generated from hydroxyacetophenone photolysis and dioxetane thermolysis, Org. Lett. 4 (2002) 537-540. [71] M.K. Hailer, P.G. Slade, B.D. Martin, K.D. Sugden. Nei deficient Escherichia coli are sensitive to chromate and accumulate the oxidized guanine lesion spiroiminodihydantoin, Chem. Res. Toxicol. 18 (2005) 1378-1383. [72] W.L. Neeley, J.M. Essigmann. Mechanisms of formation, genotoxicity, and mutation of guanine oxidation products, Chem. Res. Toxicol. 19 (2006) 491-505. [73] J. Cadet, C. Decarroz, S.Y. Wang, W.R. Midden. Mechanisms and products of photosensitized degradation of nucleic acids and related model compounds, Isr. J. Chem. 23 (1983) 420-429. [74] J. Cadet, M. Berger, C. Decarroz, J.R. Wagner, J.E. Van Lier, Y.M. Ginot, P. Vigny. Photosensitized reactions of nucleic acids, Biochimie 68 (1986) 813-834. [75] G.W. Buchko, J. Cadet, M. Berger, J.L. Ravanat. Photooxidation of d(TpG) by phthalocyanines and riboflavin. Isolation and characterization of dinucleoside monophosphates containing the 4R* and 4S* diastereoisomers of 4,8-dihydro-4- hydroxy-8-oxo-2'-deoxy- guanosine, Nucleic Acids Res. 20 (1992) 4847-4851. [76] J.-L. Ravanat, M. Berger, f. benard, R. Langlois, r. Ouellet, J.E. Van Lier, J. Cadet. Phthalocyanine and naphthalocyanine photosensitized oxidation of 2'- deoxyguanosine, Photochem. Photobiol. 55 (1992) 809-814. [77] J.-L. Ravanat, J. Cadet. Reaction of singlet oxygen with 2'-deoxyguanosine and DNA. Isolation and characterization of the main oxidation products, Chem. Res. Toxicol. 8 (1995) 379-388. [78] R.P. Hickerson, F. Prat, C.E. Muller, C.S. Foote, C.J. Burrows. Sequence and stacking dependence of 8-oxoguanine oxidation: comparison of one-electron vs singlet oxygen mechanism., J. Am. Chem. Soc. 121 (1999) 9423-9428. [79] V. Duarte, D. Gasparutto, L.F. Yamaguchi, J.-L. avanat, G.R. artinez, M.H.G. edeiros, P. i Mascio, J. adet. Oxaluric acid as the major product of singlet oxygen- mediated oxidation of 8-oxo-7,8-dihydroguanine in DNA., J. Am. Chem. Soc. 122 (2000) 12622-12628. [80] R. Misiaszek, Y. Uvaydov, C. Crean, N.E. Geacintov, V. Shafirovich. Combination reactions of superoxide with 8-Oxo-7,8-dihydroguanine radicals in DNA: kinetics and end products, J. Biol. Chem. 280 (2005) 6293-6300. [81] S.S. David, V.L. O'Shea, S. Kundu. Base-excision repair of oxidative DNA damage, Nature 447 (2007) 941-950. [82] ESCODD. Comparative analysis of baseline 8-oxo-7,8-dihydroguanine in mammalian cell DNA, by different methods in different laboratories: an approach to consensus, Carcinogenesis 23 (2002) 2129-2133. [83] ESCODD. Measurement of DNA oxidation in human cells by chromatographic and enzymic methods, Free Radic. Biol. Med. 34 (2003) 1089-1099. [84] A.R. Collins, J. Cadet, L. Moller, H.E. Poulsen, J. Vina. Are we sure we know how to measure 8-oxo-7,8-dihydroguanine in DNA from human cells?, Arch. Biochem. Biophys. 423 (2004) 57-65. [85] A. Hissung, C. von Sonntag, D. Veltwisch, K.D. Asmus. The reactions of the 2'- deoxyadenosine electron adduct in aqueous solution. The effects of the radiosensitizer p-nitroacetophenone. A pulse spectroscopic and pulse conductometric study, Int. J. Radiat. Biol. Relat Stud. Phys. Chem. Med. 39 (1981) 63-71.

63

[86] K.J. Visscher, H.J. Spoelder, H. Loman, A. Hummel, M.L. Hom. Kinetics and mechanism of electron transfer between purines and pyrimidines, their dinucleotides and polynucleotides after reaction with hydrated electrons; a pulse radiolysis study, Int. J. Radiat. Biol. 54 (1988) 787-802. [87] L.P. Candeias, S. Steenken. Electron adducts of adenine nucleosides and nucleotides in aqueous solution: Protonation at two carbon sites (C2 and C8) and intra- and intermolecular catalysis by phosphate., J. Phys. Chem. 96 (1992) 937-944. [88] H.B. Michaels, J.W. Hunt. Reactions of the hydroxyl radical with polynucleotides, Radiat. Res. 56 (1973) 57-70. [89] S. Fujita, S. Steenken. Pattern of OH radical addition to uracil and methyl- and carboxyl-substituted uracils. Electron transfer of OH adducts with N,N,N',N'- tetramethyl-p-phenylenediamine and tetranitromethane, J. Am. Chem. Soc. 103 (1981) 2540-2545. [90] Al-Sheikly, C. von Sonntag. γ−Radiolysis of 1,3-dimethyluracil in N2O-saturated aqueous solution., Z. Naturforsch. 38b (1983) 1622-1629. [91] M. Dizdaroglu, E. Holwitt, M.P. Hagan, W.F. Blakely. Formation of cytosine glycol and 5,6-dihydroxycytosine in deoxyribonucleic acid on treatment with osmium tetroxide, Biochem. J. 235 (1986) 531-536. [92] J. Cadet, T. Douki, J.L. Ravanat. Oxidatively generated base damage to cellular DNA, Free Radic. Biol. Med. 49 (2010) 9-21. [93] K. Miaskiewicz, R. Osman. Theoretical study on the deoxyribose radicals formed by hydrogen abstraction., J. Am. Chem. Soc. 116 (1994) 232-238. [94] B. Balasubramanian, W.K. Pogozelski, T.D. Tullius. DNA strand breaking by the hydroxyl radical is governed by the accessible surface areas of the hydrogen atoms of the DNA backbone, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 9738-9743. [95] W.K. Pogozelski, T.D. Tullius. Oxidative Strand Scission of Nucleic Acids: Routes Initiated by Hydrogen Abstraction from the Sugar Moiety, Chem. Rev. 98 (1998) 1089-1108. [96] M. Dizdaroglu, C. von Sonntag, D. Schulte-Frohlinde. Strand breaks and sugar release by gamma-irradiation of DNA in aqueous solution, J. Am. Chem. Soc. 97 (1975) 2277-2278. [97] M. Dizdaroglu, D. Schulte-Frohlinde, C. von Sonntag. Radiation chemistry of DNA, II. Strand breaks and sugar release by gamma-irradiation of DNA in aqueous solution. The effect of oxygen, Z. Naturforsch. 30c (1975) 826-828. [98] M. Dizdaroglu, D. Schulte-Frohlinde, C. von Sonntag. Isolation of 2-deoxy-D- erythro-pentonic acid from an alkali labile site in gamma-irradiated DNA, Int. J. Radiat. Biol. 32 (1977) 481-483. [99] M. Dizdaroglu, D. Schulte-Frohlinde, C. von Sonntag. Radiolysis of DNA in Oxygenated Aqueous Solution. Structure of An Alkali Labile Site, Z. Naturforsch. 32c (1977) 1021-1022. [100] F. Beesk, M. Dizdaroglu, D. Schulte-Frohlinde, C. von Sonntag. Radiation-induced DNA strand breaks in deoxygenated aqueous solutions. The formation of altered sugars as end groups, Int. J. Radiat. Biol. Relat Stud. Phys. Chem. Med. 36 (1979) 565-576. [101] M. Isildar, M.N. Schuchmann, D. Schulte-Frohlinde, S.C. von. gamma-Radiolysis of DNA in oxygenated aqueous solutions: alterations at the sugar moiety, Int. J. Radiat. Biol. Relat Stud. Phys. Chem. Med. 40 (1981) 347-354. [102] C. von Sonntag. Carbohydrate radicals: from ethylene glycol to DNA strand breakage, Int. J. Radiat. Biol. Relat Stud. Phys. Chem. Med. 46 (1984) 507-519.

64

[103] M. Dizdaroglu. Clemens von Sonntag and the early history of radiation-induced sugar damage in DNA, Int J Radiat. Biol. 90 (2014) 446-458. [104] K. Keck. Bildung von Cyclonucleotiden bei Betrahlung wässriger Lösungen von Purinnucleotiden, Z. Naturforsch. B 23 (1968) 1034-1043. [105] P. Jaruga, M. Dizdaroglu. 8,5'-Cyclopurine-2'-deoxynucleosides in DNA: Mechanisms of formation, measurement, repair and biological effects, DNA Repair (Amst) 7 (2008) 1413-1425. [106] J.A. Raleigh, W. Kremers, R. Whitehouse. Radiation chemistry of nucleotides: 8,5'- cyclonucleotide formation and phosphate release initiated by hydroxyl radical attack on adenosine monophosphates, Radiat. Res. 65 (1976) 414-422. [107] A.F. Fuciarelli, C.J. Koch, J.A. Raleigh. Oxygen dependence of product formation in irradiated adenosine 5'-monophosphate, Radiat. Res. 113 (1988) 447-457. [108] M. Dizdaroglu, M.L. Dirksen, H.X. Jiang, J.H. Robbins. Ionizing-radiation-induced damage in the DNA of cultured human cells. Identification of 8,5'-cyclo-2'- deoxyguanosine, Biochem. J. 241 (1987) 929-932. [109] P. Jaruga, M. Birincioglu, H. Rodriguez, M. Dizdaroglu. Mass spectrometric assays for the tandem lesion 8,5'-cyclo-2'-deoxyguanosine in mammalian DNA, Biochemistry 41 (2002) 3703-3711. [110] P. Jaruga, J. Theruvathu, M. Dizdaroglu, P.J. Brooks. Complete release of (5'S)-8,5'- cyclo-2'-deoxyadenosine from dinucleotides, oligodeoxynucleotides and DNA, and direct comparison of its levels in cellular DNA with other oxidatively induced DNA lesions, Nucleic Acids Res. 32 (2004) e87. [111] R.A. Egler, E. Fernandes, K. Rothermund, S. Sereika, N. de Souza-Pinto, P. Jaruga, M. Dizdaroglu, E.V. Prochownik. Regulation of reactive oxygen species, DNA damage, and c-Myc function by peroxiredoxin 1, Oncogene 24 (2005) 8038-8050. [112] D.C. Malins, K.M. Anderson, J.J. Stegeman, P. Jaruga, V.M. Green, N.K. Gilman, M. Dizdaroglu. Biomarkers signal contaminant effects on the organs of English sole (Parophrys vetulus) from Puget Sound, Environ. Health Perspect. 114 (2006) 823- 829. [113] K.M. Anderson, P. Jaruga, C.R. Ramsey, N.K. Gilman, V.M. Green, S.W. Rostad, J.T. Emerman, M. Dizdaroglu, D.C. Malins. Structural alterations in breast stromal and epithelial DNA: The influence of 8,5'-cyclo-2'-deoxyadenosine, Cell Cycle 5 (2006) 1240-1244. [114] M. D'errico, E. Parlanti, M. Teson, B.M. de Jesus, P. Degan, A. Calcagnile, P. Jaruga, M. Bjoras, M. Crescenzi, A.M. Pedrini, J.M. Egly, G. Zambruno, M. Stefanini, M. Dizdaroglu, E. Dogliotti. New functions of XPC in the protection of human skin cells from oxidative damage, EMBO J. 25 (2006) 4305-4315. [115] M. D'errico, E. Parlanti, M. Teson, P. Degan, T. Lemma, A. Calcagnile, I. Iavarone, P. Jaruga, M. Ropolo, A.M. Pedrini, D. Orioli, G. Frosina, G. Zambruno, M. Dizdaroglu, M. Stefanini, E. Dogliotti. The role of CSA in the response to oxidative DNA damage in human cells, Oncogene 26 (2007) 4336-4343. [116] S.G. Nyaga, P. Jaruga, A. Lohani, M. Dizdaroglu, M.K. Evans. Accumulation of oxidatively induced DNA damage in human breast cancer cell lines following treatment with hydrogen peroxide, Cell Cycle 6 (2007) 1472-1478. [117] H. Rodriguez, P. Jaruga, D. Leber, S.G. Nyaga, M.K. Evans, M. Dizdaroglu. Lymphoblasts of women with BRCA1 mutations are deficient in cellular repair of 8,5'-Cyclopurine-2'-deoxynucleosides and 8-hydroxy-2'-deoxyguanosine, Biochemistry 46 (2007) 2488-2496.

65

[118] G. Kirkali, M. Tunca, S. Genc, P. Jaruga, M. Dizdaroglu. Oxidative DNA damage in polymorphonuclear leukocytes of patients with familial Mediterranean fever, Free Radic. Biol. Med. 44 (2008) 386-393. [119] G. Gokce, G. Ozsarlak-Sozer, G. Oktay, G. Kirkali, P. Jaruga, M. Dizdaroglu, Z. Kerry. Glutathione depletion by buthionine sulfoximine induces oxidative damage to DNA in organs of rabbits in vivo., Biochemistry 48 (2009) 4980-4987. [120] G. Kirkali, N.C. de Souza-Pinto, P. Jaruga, V.A. Bohr, M. Dizdaroglu. Accumulation of (5'S)-8,5'-cyclo-2'-deoxyadenosine in organs of Cockayne syndrome complementation group B gene knockout mice, DNA Repair (Amst) 8 (2009) 274-278. [121] P. Jaruga, Y. Xiao, B.C. Nelson, M. Dizdaroglu. Measurement of (5'R)- and (5'S)- 8,5'-cyclo-2'-deoxyadenosines in DNA in vivo by liquid chromatography/isotope- dilution tandem mass spectrometry, Biochem. Biophys. Res. Commun. 386 (2009) 656-660. [122] P. Jaruga, Y. Xiao, V. Vartanian, R.S. Lloyd, M. Dizdaroglu. Evidence for the involvement of DNA repair enzyme NEIL1 in nucleotide excision repair of (5'R)- and (5'S)-8,5'-cyclo-2'-deoxyadenosines, Biochemistry 49 (2010) 1053-1055. [123] G. Kirkali, D. Keles, A.E. Canda, C. Terzi, P.T. Reddy, P. Jaruga, M. Dizdaroglu, G. Oktay. Evidence for upregulated repair of oxidatively induced DNA damage in human colorectal cancer, DNA Repair (Amst) 10 (2011) 1114-1120. [124] J. Wang, B. Yuan, C. Guerrero, R. Bahde, S. Gupta, Y. Wang. Quantification of oxidative DNA lesions in tissues of Long-Evans Cinnamon rats by capillary high- performance liquid chromatography-tandem mass spectrometry coupled with stable isotope-dilution method, Anal. Chem. 83 (2011) 2201-2209. [125] J.S. Tilstra, A.R. Robinson, J. Wang, S.Q. Gregg, C.L. Clauson, D.P. Reay, L.A. Nasto, C.M. St Croix, A. Usas, N. Vo, J. Huard, P.R. Clemens, D.B. Stolz, D.C. Guttridge, S.C. Watkins, G.A. Garinis, Y. Wang, L.J. Niedernhofer, P.D. Robbins. NF-kappaB inhibition delays DNA damage-induced senescence and aging in mice, J. Clin. Invest. 122 (2012) 2601-2612. [126] J. Wang, C.L. Clauson, P.D. Robbins, L.J. Niedernhofer, Y. Wang. The oxidative DNA lesions 8,5'-cyclopurines accumulate with aging in a tissue-specific manner, Aging Cell 11 (2012) 714-716. [127] P. Jaruga, M. Dizdaroglu. Identification and quantification of (5'R)- and (5'S)-8,5'- cyclo-2'-deoxyadenosines in human urine as putative biomarkers of oxidatively induced damage to DNA, Biochem. Biophys. Res. Commun. 397 (2010) 48-52. [128] P. Jaruga, R. Rozalski, A. Jawien, A. Migdalski, R. Olinski, M. Dizdaroglu. DNA damage products (5'R)- and (5'S)-8,5'-cyclo-2'-deoxyadenosines as potential biomarkers in human urine for atherosclerosis, Biochemistry 51 (2012) 1822-1824. [129] N. Belmadoui, F. Boussicault, M. Guerra, J.L. Ravanat, C. Chatgilialoglu, J. Cadet. Radiation-induced formation of purine 5',8-cyclonucleosides in isolated and cellular DNA: high stereospecificity and modulating effect of oxygen, Org. Biomol. Chem. 8 (2010) 3211-3219. [130] C. Chatgilialoglu, C. Ferreri, M.A. Terzidis. Purine 5',8-cyclonucleoside lesions: chemistry and biology, Chem. Soc. Rev. 40 (2011) 1368-1382. [131] J. Cadet, J.R. Wagner. DNA base damage by reactive oxygen species, oxidizing agents, and UV radiation, in: E.C. Friedberg, S.J. Elledge, A.R. Lehmann, T. Lindahl and M. Muzi-Falconi (Eds.), DNA Repair, Mutagenesis, and Other responses to DNA Damage, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, 2014, pp. 1-18.

66

[132] H.C. Box, E.E. Budzinski, H.G. Freund, M.S. Evans, H.B. Patrzyc, J.C. Wallace, A.E. MacCubbin. Vicinal lesions in X-irradiated DNA?, Int. J. Radiat. Biol. 64 (1993) 261-263. [133] E.E. Budzinski, A.E. MacCubbin, H.G. Freund, J.C. Wallace, H.C. Box. Characterization of the products of dinucleoside monophosphates d(GpN) irradiated in aqueous solutions, Radiat. Res. 136 (1993) 171-177. [134] E.E. Budzinski, J.D. Dawidzik, J.C. Wallace, H.G. Freund, H.C. Box. The radiation chemistry of d(CpGpTpA) in the presence of oxygen, Radiat. Res. 142 (1995) 107- 109. [135] E. Schroder, E.E. Budzinski, J.C. Wallace, J.D. Zimbrick, H.C. Box. Radiation chemistry of d(ApCpGpT), Int. J. Radiat. Biol. 68 (1995) 509-523. [136] H.C. Box, H.G. Freund, E.E. Budzinski, J.C. Wallace, A.E. MacCubbin. Free radical-induced double base lesions, Radiat. Res. 141 (1995) 91-94. [137] H.C. Box, E.E. Budzinski, J.B. Dawidzik, J.S. Gobey, H.G. Freund. Free radical- induced tandem base damage in DNA oligomers, Free Radic. Biol. Med. 23 (1997) 1021-1030. [138] H.C. Box, H.B. Patrzyc, J.B. Dawidzik, J.C. Wallace, H.G. Freund, H. Iijima, E.E. Budzinski. Double base lesions in DNA X-irradiated in the presence or absence of oxygen, Radiat. Res. 153 (2000) 442-446. [139] A.E. MacCubbin, H. Iijima, N. Ersing, J.B. Dawidzik, H.B. Patrzyc, J.C. Wallace, E.E. Budzinski, H.G. Freund, H.C. Box. Double-base lesions are produced in DNA by free radicals, Arch. Biochem. Biophys. 375 (2000) 119-123. [140] H.B. Patrzyc, J.B. Dawidzik, E.E. Budzinski, H. Iijima, H.C. Box. Double lesions are produced in DNA oligomer by ionizing radiation and by metal-catalyzed H2O2 reactions, Radiat. Res. 155 (2001) 634-636. [141] A.-G. Bourdat, T. Douki, S. Frelon, D. Gasparutto, J. Cadet. Tandem base lesions are generated by hydroxyl radical within isolated DNA in aerated aqueous solution., J. Am. Chem. Soc. 122 (2000) 4549-4566. [142] T. Douki, J. Riviere, J. Cadet. DNA tandem lesions containing 8-oxo-7,8- dihydroguanine and formamido residues arise from intramolecular addition of thymine peroxyl radical to guanine, Chem. Res. Toxicol. 15 (2002) 445-454. [143] J. Cadet, S. Bellon, M. Berger, A.G. Bourdat, T. Douki, V. Duarte, S. Frelon, D. Gasparutto, E. Muller, J.L. Ravanat, S. Sauvaigo. Recent aspects of oxidative DNA damage: guanine lesions, measurement and substrate specificity of DNA repair glycosylases, Biol. Chem. 383 (2002) 933-943. [144] A. Romieu, S. Bellon, D. Gasparutto, J. Cadet. Synthesis and UV photolysis of oligodeoxynucleotides that contain 5- (phenylthiomethyl)-2'-deoxyuridine: a specific photolabile precursor of 5-(2'-deoxyuridilyl)methyl radical, Org. Lett. 2 (2000) 1085-1088. [145] S. Bellon, J.L. Ravanat, D. Gasparutto, J. Cadet. Cross-linked thymine-purine base tandem lesions: synthesis, characterization, and measurement in gamma-irradiated isolated DNA, Chem. Res. Toxicol. 15 (2002) 598-606. [146] H. Hong, H. Cao, Y. Wang, Y. Wang. Identification and quantification of a guanine-thymine intrastrand cross-link lesion induced by Cu(II)/H2O2/ascorbate, Chem. Res. Toxicol. 19 (2006) 614-621. [147] S. Bellon, D. Gasparutto, C. Saint-Pierre, J. Cadet. Guanine-thymine intrastrand cross-linked lesion containing oligonucleotides: from chemical synthesis to in vitro enzymatic replication, Org. Biomol. Chem. 4 (2006) 3831-3837.

67

[148] V. Labet, C. Morell, A. Grand, J. Cadet, P. Cimino, V. Barone. Formation of cross- linked adducts between guanine and thymine mediated by hydroxyl radical and one- electron oxidation: a theoretical study, Org. Biomol. Chem. 6 (2008) 3300-3305. [149] Y. Wang. Bulky DNA lesions induced by reactive oxygen species., Chem. Res. Toxicol. 21 (2008) 276-281. [150] B. Xerri, C. Morell, A. Grand, J. Cadet, P. Cimino, V. Barone. Radiation-induced formation of DNA intrastrand crosslinks between thymine and adenine bases: A theoretical approach, Org. Biomol. Chem. 4 (2006) 3986-3992. [151] C. Gu, Y. Wang. LC-MS/MS identification and yeast polymerase eta bypass of a novel gamma-irradiation-induced intrastrand cross-link lesion G[8-5]C, Biochemistry 43 (2004) 6745-6750. [152] Q. Zhang, Y. Wang. Independent generation of 5-(2'-deoxycytidyl)methyl radical and the formation of a novel crosslink lesion between 5-methylcytosine and guanine, J. Am. Chem. Soc. 125 (2003) 12795-12802. [153] Q. Zang, Y. Wang. Generation of 5-(2'-deoxycytidyl)methyl radical and the formation of intrastrand cross-link lesions in oligodeoxyribonucleosides., Nucleic Acids Res. 33 (2005) 1593-1603. [154] H. Cao, Y. Wang. Quantification of oxidative single-base and intrastrand cross-link lesions in unmethylated and CpG-methylated DNA induced by Fenton-type reagents, Nucleic Acids Res. 35 (2007) 4833-4844. [155] I.S. Hong, M.M. Greenberg. Efficient DNA interstrand cross-link formation from a nucleotide radical, J. Am. Chem. Soc. 127 (2005) 3692-3693. [156] I.S. Hong, H. Ding, M.M. Greenberg. Oxygen independent DNA interstrand cross- link formation by a nucleotide radical, J. Am. Chem. Soc. 128 (2006) 485-491. [157] H. Ding, M.M. Greenberg. Gamma-radiolysis and hydroxyl radical produce interstrand cross-links in DNA involving , Chem. Res. Toxicol. 20 (2007) 1623-1628. [158] H. Dink, A. Majumdar, J.R. Tolman, M.M. Greenberg. Multinuclear NMR and kinetic analysis of DNA interstrand cross-link formation., J. Am. Chem. Soc. 130 (2008) 17981-17987. [159] Y. Jiang, H. Hong, H. Cao, Y. Wang. In vivo formation and in vitro replication of a guanine-thymine intrastrand cross-link lesion, Biochemistry 46 (2007) 12757- 12763. [160] H. Hong, H. Cao, Y. Wang. Formation and genotoxicity of a guanine-cytosine intrastrand cross-link lesion in vivo, Nucleic Acids Res. 35 (2007) 7118-7127. [161] J. Wang, H. Cao, C. You, B. Yuan, R. Bahde, S. Gupta, C. Nishigori, L.J. Niedernhofer, P.J. Brooks, Y. Wang. Endogenous formation and repair of oxidatively induced G[8-5 m]T intrastrand cross-link lesion, Nucleic Acids Res. 40 (2012) 7368-7374. [162] O. Yamamoto. Ionizing radiation-induced DNA-protein cross-linking., in: K.C. Smith (Ed.), Aging, Carcinogenesis, and Radiation Biology, Plenum Press, New York, 1976, pp. 165-192. [163] A.J. Fornace, Jr., J.B. Little. DNA-protein cross-linking by chemical carcinogens in mammalian cells, Cancer Res. 39 (1979) 704-710. [164] L.K. Mee, S.J. Adelstein. Radiolysis of chromatin extracted from cultured mammalian cells: formation of DNA-protein cross links, Int. J. Radiat. Biol. 36 (1979) 359-366. [165] L.K. Mee, S.J. Adelstein. Predominance of core histones in formation of DNA- protein cross- links in g-irradiated chromatin, Proc. Natl. Acad. Sci. U. S. A. 78 (1981) 2194-2198.

68

[166] S.A. Lesko, J.L. Drocourt, S.U. Yang. Deoxyribonucleic acid- protein and deoxyribonucleic acid interstrand cross-links induced in isolated chromatin byhydrogen peroxide and ferrous ethylenediaminetetraacetate chelates, Biochemistry 21 (1982) 5010-5015. [167] A.E. Cress, G.T. Bowden. Covalent DNA-protein cross-linking occurs after hyperthermia and radiation, Radiat. Res. 95 (1983) 610-618. [168] N.L. Oleinick, S. Chiu, N. Ramakrishnan, L. Xue. The formation, identification,and significance of DNA-protein cross-links in mammalian cells, Brit. J. Cancer 55 (Suppl.VIII) (1987) 135-140. [169] M. Dizdaroglu. The use of capillary gas chromatography-mass spectrometry for identification of radiation-induced DNA base damage and DNA base-amino acid crosslinks, J. Chromatogr. 295 (1984) 103-121. [170] M.G. Simic, M. Dizdaroglu. Formation of Radiation-Induced Crosslinks between Thymine and Tyrosine: Possible Model for Crosslinking of DNA and Proteins by Ionizing Radiation, Biochemistry 24 (1985) 233-236. [171] S. Margolis, B. Coxon, E. Gajewski, M. Dizdaroglu. Structure of a hydroxyl radical induced cross-link of thymine and tyrosine, Biochemistry 27 (1988) 6353-6359. [172] M.S.W. Lipton, A.F. Fuciarelli, D.L. Springer, C.G. Edmonds. Characterization of radiation-induced thymine-tyrosine crosslinks by electrospray ionization mass spectrometry, Radiat. Res. 145 (1996) 681-686. [173] M.S. Lipton, A.L. Fuciarelli, D.L. Springer, S.A. Hofstadler, C.G. Edmonds. Analysis of radiation induced nucleobase-peptide crosslinks by electrospray ionization mass spectrometry, Rapid Commun. Mass Spectrom. 11 (1997) 1673- 1676. [174] T.S. Carlton, B.A. Ingelse, D.S. Black, D.C. Craig, K.E. Mason, M.W. Duncan. A covalent thymine-tyrosine adduct involved in DNA-protein crosslinks: Synthesis, characterization and quantification, Free Radic. Biol. Med. 27 (1999) 254-261. [175] M. Dizdaroglu, E. Gajewski. Structure and mechanism of hydroxyl radical-induced formation of a DNA- protein cross-link involving thymine and lysine in nucleohistone, Cancer Res. 49 (1989) 3463-3467. [176] Z. Nackerdien, G. Rao, M.A. Cacciuttolo, E. Gajewski, M. Dizdaroglu. Chemical nature of DNA-protein cross-links produced in mammalian chromatin by hydrogen peroxide in the presence of iron or copper ions, Biochemistry 30 (1991) 4873-4879. [177] E.J. Land, M. Ebert. Pulse radiolysis studies of aqueous phenol. Water elimination from dihydroxycyclohexadienyl radicals to form phenoxyl, Trans. Farad. Soc. 63 (1967) 1181-1190. [178] R. Olinski, Z. Nackerdien, M. Dizdaroglu. DNA-protein cross-linking between thymine and tyrosine in chromatin of gamma-irradiated or H2O2-treated cultured human cells., Archiv. Biochem. Biophys. 297 (1992) 139-143. [179] S.A. Altman, T.H. Zastawny, L. Randers-Eichhorn, M.A. Cacciuttolo, S.A. Akman, M. Dizdaroglu, G. Rao. Formation of DNA-protein cross-links in cultured mammalian cells upon treatment with iron ions, Free Radic. Biol. Med. 19 (1995) 897-902. [180] S. Toyokuni, T. Mori, H. Hiai, M. Dizdaroglu. Treatment of Wistar rats with a renal carcinogen, ferric nitrilotriacetate, causes DNA-protein cross-linking between thymine and tyrosine in their renal chromatin, Int. J. Cancer 62 (1995) 309-313. [181] E. Gajewski, A. Fuciarelli, M. Dizdaroglu. Structure of hydroxyl radical-induced DNA-protein crosslinks in calf thymus nucleohistone in vitro, Int. J. Radiat. Biol. 54 (1988) 445-459.

69

[182] E. Gajewski, M. Dizdaroglu. Hydroxyl Radical-Induced Cross-linking of Cytosine and Tyrosine in Nucleohistone, Biochemistry 29 (1990) 977-980. [183] S. Perrier, J. Hau, D. Gasparutto, J. Cadet, A. Favier, J.L. Ravanat. Characterization of Lysine-Guanine Cross-Links upon One-Electron Oxidation of a Guanine- Containing Oligonucleotide in the Presence of a Trilysine Peptide, J. Am. Chem. Soc. 128 (2006) 5703-5710. [184] J.F. Ward. Some biochemical consequences of the spatial distribution of ionizing radiation-produced free radicals, Radiat. Res. 86 (1981) 185-195. [185] J.F. Ward. Biochemistry of DNA lesions, Radiat. Res. Suppl 8 (1985) S103-S111. [186] D.T. Goodhead. Initial events in the cellular effects of ionizing radiations: clustered damage in DNA, Int. J. Radiat. Biol. 65 (1994) 7-17. [187] J.F. Ward. The complexity of DNA damage: relevance to biological consequences, Int. J. Radiat. Biol. 66 (1994) 427-432. [188] B.M. Sutherland, P.V. Bennett, O. Sidorkina, J. Laval. Clustered DNA damages induced in isolated DNA and in human cells by low doses of ionizing radiation, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 103-108. [189] B.M. Sutherland, P.V. Bennett, J.C. Sutherland, J. Laval. Clustered DNA damages induced by x rays in human cells, Radiat. Res. 157 (2002) 611-616. [190] A.G. Georgakilas. Processing of DNA damage clusters in human cells: current status of knowledge, Mol. Biosyst. 4 (2008) 30-35. [191] J.H. Hoeijmakers. Genome maintenance mechanisms for preventing cancer, Nature 411 (2001) 366-374. [192] R.D. Wood, M. Mitchell, T. Lindahl. Human DNA repair genes, Mutat. Res. 577 (2005) 275-283. [193] M.M. Slupska, C. Baikalov, W.M. Luther, J.H. Chiang, Y.F. Wei, J.H. Miller. Cloning and sequencing a human homolog (hMYH) of the Escherichia coli mutY gene whose function is required for the repair of oxidative DNA damage, J. Bacteriol. 178 (1996) 3885-3892. [194] M. Takao, Q.M. Zhang, S. Yonei, A. Yasui. Differential subcellular localization of human MutY homolog (hMYH) and the functional activity of adenine:8-oxoguanine DNA glycosylase, Nucleic Acids Res. 27 (1999) 3638-3644. [195] T. Ohtsubo, K. Nishioka, Y. Imaiso, S. Iwai, H. Shimokawa, H. Oda, T. Fujiwara, Y. Nakabeppu. Identification of human MutY homolog (hMYH) as a repair enzyme for 2-hydroxyadenine in DNA and detection of multiple forms of hMYH located in nuclei and mitochondria, Nucleic Acids Res. 28 (2000) 1355-1364. [196] A.R. Parker, J.R. Eshleman. Human MutY: gene structure, protein functions and interactions, and role in carcinogenesis, Cell Mol. Life Sci. 60 (2003) 2064-2083. [197] H. Maki, M. Sekiguchi. MutT protein specifically hydrolyses a potent mutagenic substrate for DNA synthesis, Nature 355 (1992) 273-275. [198] J.Y. Mo, H. Maki, M. Sekiguchi. Hydrolytic elimination of a mutagenic nucleotide, 8-oxodGTP, by human 18-kilodalton protein: sanitization of nucleotide pool, Proc. Natl. Acad. Sci. U. S. A. 89 (1992) 11021-11025. [199] K. Sakumi, M. Furuichi, T. Tsuzuki, T. Kakuma, S. Kawabata, H. Maki, M. Sekiguchi. Cloning and expression of cDNA for a human enzyme that hydrolyzes 8- oxo-dGTP, a mutagenic substrate for DNA synthesis, J. Biol. Chem. 268 (1993) 23524-23530. [200] T. Tsuzuki, A. Egashira, H. Igarashi, T. Iwakuma, Y. Nakatsuru, Y. Tominaga, H. Kawate, K. Nakao, K. Nakamura, F. Ide, S. Kura, Y. Nakabeppu, M. Katsuki, T. Ishikawa, M. Sekiguchi. Spontaneous tumorigenesis in mice defective in the MTH1

70

gene encoding 8-oxo-dGTPase, Proc. Natl. Acad. Sci. U. S. A 98 (2001) 11456- 11461. [201] M. Sekiguchi, T. Tsuzuki. Oxidative nucleotide damage: consequences and prevention, Oncogene 21 (2002) 8895-8904. [202] Y. Sakai, M. Furuichi, M. Takahashi, M. Mishima, S. Iwai, M. Shirakawa, Y. Nakabeppu. A molecular basis for the selective recognition of 2-hydroxy-dATP and 8-oxo-dGTP by human MTH1, J. Biol. Chem. 277 (2002) 8579-8587. [203] E. Speina, K.D. Arczewska, D. Gackowski, M. Zielinska, A. Siomek, J. Kowalewski, R. Olinski, B. Tudek, J.T. Kusmierek. Contribution of hMTH1 to the maintenance of 8-oxoguanine levels in lung DNA of non-small-cell lung cancer patients, J Natl. Cancer Inst. 97 (2005) 384-395. [204] H. Gad, T. Koolmeister, A.S. Jemth, S. Eshtad, S.A. Jacques, C.E. Strom, L.M. Svensson, N. Schultz, T. Lundback, B.O. Einarsdottir, A. Saleh, C. Gokturk, P. Baranczewski, R. Svensson, R.P. Berntsson, R. Gustafsson, K. Stromberg, K. Sanjiv, M.C. Jacques-Cordonnier, M. Desroses, A.L. Gustavsson, R. Olofsson, F. Johansson, E.J. Homan, O. Loseva, L. Brautigam, L. Johansson, A. Hoglund, A. Hagenkort, T. Pham, M. Altun, F.Z. Gaugaz, S. Vikingsson, B. Evers, M. Henriksson, K.S. Vallin, O.A. Wallner, L.G. Hammarstrom, E. Wiita, I. Almlof, C. Kalderen, H. Axelsson, T. Djureinovic, J.C. Puigvert, M. Haggblad, F. Jeppsson, U. Martens, C. Lundin, B. Lundgren, I. Granelli, A.J. Jensen, P. Artursson, J.A. Nilsson, P. Stenmark, M. Scobie, U.W. Berglund, T. Helleday. MTH1 inhibition eradicates cancer by preventing sanitation of the dNTP pool, Nature 508 (2014) 215-221. [205] J. Rouse, S.P. Jackson. Interfaces between the detection, signaling, and repair of DNA damage, Science 297 (2002) 547-551. [206] S.P. Jackson. Sensing and repairing DNA double-strand breaks, Carcinogenesis 23 (2002) 687-696. [207] P. Fortini, E. Parlanti, O.M. Sidorkina, J. Laval, E. Dogliotti. The type of DNA glycosylase determines the base excision repair pathway in mammalian cells, J. Biol. Chem. 274 (1999) 15230-15236. [208] A.K. McCullough, M.L. Dodson, R.S. Lloyd. Initiation of base excision repair: glycosylase mechanisms and structures, Annu. Rev. Biochem. 68 (1999) 255-285. [209] A. Sancar, L.A. Lindsey-Boltz, K. Unsal-Kacmaz, S. Linn. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints, Annu. Rev. Biochem. 73 (2004) 39-85. [210] J.A. Eisen, P.C. Hanawalt. A phylogenomic study of DNA repair genes, proteins, and processes, Mutat. Res. 435 (1999) 171-213. [211] L. Aravind, D.R. Walker, E.V. Koonin. Conserved domains in DNA repair proteins and evolution of repair systems, Nucleic Acids Res. 27 (1999) 1223-1242. [212] A. Prakash, S. Doublie, S.S. Wallace. The Fpg/Nei family of DNA glycosylases: substrates, structures, and search for damage, Prog. Mol. Biol. Transl. Sci. 110 (2012) 71-91. [213] S.S. David, S.D. Williams. Chemistry of glycosylases and endonucleases involved in base-excision repair, Chem. Rev. 98 (1998) 1221-1262. [214] H. Sampath, A.K. McCullough, R.S. Lloyd. Regulation of DNA glycosylases and their role in limiting disease, Free Radic. Res. 46 (2012) 460-478. [215] S.D. Kathe, R. Barrantes-Reynolds, P. Jaruga, M.R. Newton, C.J. Burrows, V. Bandaru, M. Dizdaroglu, J.P. Bond, S.S. Wallace. Plant and fungal Fpg homologs are formamidopyrimidine DNA glycosylases but not 8-oxoguanine DNA glycosylases, DNA Repair (Amst) 8 (2009) 643-653.

71

[216] M. Dizdaroglu. Substrate specificities and excision kinetics of DNA glycosylases involved in base-excision repair of oxidative DNA damage, Mutat. Res. 531 (2003) 109-126. [217] M. Dizdaroglu. Base-excision repair of oxidative DNA damage by DNA glycosylases, Mutat. Res. 591 (2005) 45-59. [218] C.J. Chetsanga, T. Lindahl. Release of 7-methylguanine residues whose imidazole rings have been opened from damaged DNA by a DNA glycosylase from Escherichia coli, Nucleic Acids Res. 6 (1979) 3673-3684. [219] C.J. Chetsanga, M. Lozon, C. Makaroff, L. Savage. Purification and characterization of Escherichia coli formamidopyrimidine-DNA glycosylase that excises damaged 7-methylguanine from deoxyribonucleic acid, Biochemistry 20 (1981) 5201-5207. [220] L. Breimer. Enzymatic excision form gamma-irradiated polydeoxyribonucleotides of adenine residues whose imidazole rings have been ruptured, Nucleic Acids Res. 12 (1984) 6359-6367. [221] J. Tchou, H. Kasai, S. Shibutani, M.-H. Chung, J. Laval, A.P. Grollman, S. Nishimura. 8-oxoguanine (8-hydroxyguanine) DNA glycosylase and its substrate specificity, Proc. Natl. Acad. Sci. U. S. A. 88 (1991) 4690-4694. [222] S.S. Wallace. Enzymatic processing of radiation-induced free radical damage in DNA, Radiat. Res. 150 (1998) S60-S79. [223] A. Karakaya, P. Jaruga, V.A. Bohr, A.P. Grollman, M. Dizdaroglu. Kinetics of excision of purine lesions from DNA by Escherichia coli Fpg protein, Nucleic Acids Res. 25 (1997) 474-479. [224] B. Castaing, A. Geiger, H. Seliger, P. Nehls, J. Laval, C. Zelwer, S. Boiteux. Cleavage and binding of a DNA fragment containing a single 8-oxoguanine by wild type and mutant FPG proteins, Nucleic Acids Res. 21 (1993) 2899-2905. [225] J. Tchou, V. Bodepudi, S. Shibutani, I. Antoshechkin, J. Miller, A.P. Grollman, F. Johnson. Substrate specificity of Fpg protein. Recognition and cleavage of oxidatively damaged DNA, J. Biol. Chem. 269 (1994) 15318-15324. [226] O. Sidorkina, M. Dizdaroglu, J. Laval. Effect of single mutations on the specificity of Escherichia coli FPG protein for excision of purine lesions from DNA damaged by free radicals, Free Radic. Biol. Med. 31 (2001) 816-823. [227] J.R. Battista. Against all odds: the survival strategies of Deinococcus radiodurans, Annu. Rev. Microbiol. 51 (1997) 203-224. [228] M.S. Lipton, L. Pasa-Tolic', G.A. Anderson, D.J. Anderson, D.L. Auberry, J.R. Battista, M.J. Daly, J. Fredrickson, K.K. Hixson, H. Kostandarithes, C. Masselon, L.M. Markillie, R.J. Moore, M.F. Romine, Y. Shen, E. Stritmatter, N. Tolic', H.R. Udseth, A. Venkateswaran, K.K. Wong, R. Zhao, R.D. Smith. Global analysis of the Deinococcus radiodurans proteome by using accurate mass tags, Proc. Natl. Acad. Sci. U. S. A 99 (2002) 11049-11054. [229] S. Sentürker, C. Bauche, J. Laval, M. Dizdaroglu. Substrate specificity of Deinococcus radiodurans Fpg protein, Biochemistry 38 (1999) 9435-9439. [230] M. Radman. An endonuclease from Escherichia coli that introduces single polynucleotide chain scissions in ultraviolet-irradiated DNA, J. Biol. Chem. 251 (1976) 1438-1445. [231] R.P. Cunningham, B. Weiss. Endonuclease III (nth) mutants of Escherichia coli, Proc. Natl. Acad. Sci. U. S. A 82 (1985) 474-478. [232] H. Asahara, P.M. Wistort, J.F. Bank, R.H. Bakerian, R.P. Cunningham. Purification and characterization of Escherichia coli endonuclease III from the cloned nth gene, Biochemistry 28 (1989) 4444-4449.

72

[233] R.P. Cunningham, H. Asahara, J.F. Bank, C.P. Scholes, J.C. Salerno, K. Surerus, E. Munck, J. McCracken, J. Peisach, M.H. Emptage. Endonuclease III is an iron-sulfur protein, Biochemistry 28 (1989) 4450-4455. [234] L. Gros, M.K. Saparbaev, J. Laval. Enzymology of the repair of free radicals- induced DNA damage, Oncogene 21 (2002) 8905-8925. [235] M. Dizdaroglu, J. Laval, S. Boiteux. Substrate specificity of Escherischia coli endonuclease III: excision of thymine- and cytosine-derived lesions in DNA produced by ionizing radiation-generated free radicals, Biochemistry 32 (1993) 12105-12111. [236] M. Dizdaroglu, C. Bauche, H. Rodriguez, J. Laval. Novel substrates of Escherichia coli Nth protein and its kinetics for excision of modified bases from DNA damaged by free radicals, Biochemistry 39 (2000) 5586-5592. [237] R.J. Melamede, Z. Hatahet, Y.W. Kow, H. Ide, S.S. Wallace. Isolation and characterization of endonuclease VIII from Escherichia coli, Biochemistry 33 (1994) 1255-1264. [238] D. Jiang, Z. Hatahet, J.O. Blaisdell, R.J. Melamede, S.S. Wallace. Escherichia coli endonuclease VIII: cloning, sequencing, and overexpression of the nei structural gene and characterization of nei and nei nth mutants, J. Bacteriol. 179 (1997) 3773- 3782. [239] D. Jiang, Z. Hatahet, R.J. Melamede, Y.W. Kow, S.S. Wallace. Characterization of Escherichia coli endonuclease VIII, J. Biol. Chem. 272 (1997) 32230-32239. [240] S. Burgess, P. Jaruga, M.L. Dodson, M. Dizdaroglu, R.S. Lloyd. Determination of active site residues in Escherichia coli endonuclease VIII, J. Biol. Chem. 277 (2002) 2938-2944. [241] M. Dizdaroglu, S.M. Burgess, P. Jaruga, T.K. Hazra, H. Rodriguez, R.S. Lloyd. Substrate specificity and excision kinetics of Escherichia coli endonuclease VIII (Nei) for modified bases in DNA damaged by free radicals, Biochemistry 40 (2001) 12150-12156. [242] T. Lindahl, S. Ljungquist, W. Siegert, B. Nyberg, B. Sperens. DNA N-glycosidases: properties of uracil-DNA glycosidase from Escherichia coli, J Biol. Chem. 252 (1977) 3286-3294. [243] Z. Hatahet, Y.W. Kow, A.A. Purmal, R.P. Cunningham, S.S. Wallace. New substrates for old enzymes. 5-Hydroxy-2'-deoxycytidine and 5-hydroxy-2'- deoxyuridine are substrates for Escherichia coli endonuclease III and formamidopyrimidine DNA N-glycosylase, while 5- hydroxy-2'-deoxyuridine is a substrate for uracil DNA N-glycosylase, J. Biol. Chem. 269 (1994) 18814-18820. [244] T.H. Zastawny, P.W. Doetsch, M. Dizdaroglu. A novel activity of E. coli uracil DNA N-glycosylase excision of isodialuric acid (5,6-dihydroxyuracil), a major product of oxidative DNA damage, from DNA, FEBS Lett. 364 (1995) 255-258. [245] J.P. McGoldrick, Y.-C. Yeh, M. Solomon, J.M. Essigmann, A.-L. Lu. Characterization of a mammalian homolog of the Escherichia coli MutY mismatch repair protein, Mol. Cell. Biol. 15 (1995) 989-996. [246] T. Roldán-Arjona, C. Anselmino, T. Lindahl. Molecular cloning and functional analysis of a Schizosaccharomyces pombe homologue of Escherichia coli endonuclease III, Nucleic Acids Res. 24 (1996) 3307-3312. [247] R. Aspinwall, D.G. Rothwell, T. Roldan-Arjona, C. Anselmino, C.J. Ward, J.P. Cheadle, J.R. Sampson, T. Lindahl, P.C. Harris, I.D. Hickson. Cloning and characterization of a functional human homolog of Escherichia coli endonuclease III, Proc. Natl. Acad. Sci. USA 94 (1997) 109-114.

73

[248] B. Karahalil, T. Roldan-Arjona, M. Dizdaroglu. Substrate specificity of Schizosaccharomyces pombe Nth protein for products of oxidative DNA damage, Biochemistry 37 (1998) 590-595. [249] M. Dizdaroglu, B. Karahalil, S. Sentürker, T.J. Buckley, T. Roldan-Arjona. Excision of products of oxidative DNA base damage by human NTH1 protein, Biochemistry 38 (1999) 243-246. [250] J. Hu, N.C. de Souza-Pinto, K. Haraguchi, B.A. Hogue, P. Jaruga, M.M. Greenberg, M. Dizdaroglu, V.A. Bohr. Repair of formamidopyrimidines in DNA involves different glycosylases: role of the OGG1, NTH1, and NEIL1 enzymes, J. Biol. Chem. 280 (2005) 40544-40551. [251] M.K. Chan, M.T. Ocampo-Hafalla, V. Vartanian, P. Jaruga, G. Kirkali, K.L. Koenig, S. Brown, R.S. Lloyd, M. Dizdaroglu, G.W. Teebor. Targeted deletion of the genes encoding NTH1 and NEIL1 DNA N-glycosylases reveals the existence of novel carcinogenic oxidative damage to DNA, DNA Repair (Amst) 8 (2009) 786- 794. [252] C. Dherin, J.P. Radicella, M. Dizdaroglu, S. Boiteux. Excision of oxidatively damaged DNA bases by the human alpha-hOgg1 protein and the polymorphic alpha-hOgg1(Ser326Cys) protein which is frequently found in human populations, Nucleic Acids Res. 27 (1999) 4001-4007. [253] R.H. Elder, G.L. Dianov. Repair of dihydrouracil supported by base excision repair in mNTH1 knock-out cell extracts, J Biol. Chem. 277 (2002) 50487-50490. [254] L. Eide, M. Bjorås, M. Pirovano, I. Alseth, K.G. Berdal, E. Seeberg. Base excision of oxidative purine and pyrimidine DNA damage in Saccharomyces cerevisiae by a DNA glycosylase with sequence similarity to endonuclease III from Escherichia coli, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 10735-10740. [255] L. Augeri, Y.M. Lee, A.B. Barton, P.W. Doetsch. Purification, characterization, gene cloning, and expression of Saccharomyces cerevisiae redoxyendonuclease, a homolog of Escherichia coli endonuclease III, Biochemistry 36 (1997) 721-729. [256] H.J. You, R.L. Swanson, P.W. Doetsch. Saccharomyces cerevisiae possesses two functional homologues of Escherichia coli endonuclease III, Biochemistry 37 (1998) 6033-6040. [257] H.J. You, R.L. Swanson, C. Harrington, A.H. Corbett, S. Jinks-Robertson, S. Senturker, S.S. Wallace, S. Boiteux, M. Dizdaroglu, P.W. Doetsch. Saccharomyces cerevisiae Ntg1p and Ntg2p: broad specificity N- glycosylases for the repair of oxidative DNA damage in the nucleus and mitochondria, Biochemistry 38 (1999) 11298-11306. [258] S. Sentürker, P. Auffret van der Kemp, H.J. You, P.W. Doetsch, M. Dizdaroglu, S. Boiteux. Substrate specificities of the Ntg1 and Ntg2 proteins of Saccharomyces cerevisiae for oxidized DNA bases are not identical, Nucleic Acids Res. 26 (1998) 5270-5276. [259] G. Slupphaug, I. Eftedal, B. Kavli, S. Bharati, N.M. Helle, T. Haug, D.W. Levine, H.E. Krokan. Properties of a recombinant human uracil-DNA glycosylase from the UNG gene and evidence that UNG encodes the major uracil-DNA glycosylase, Biochemistry 34 (1995) 128-138. [260] M. Dizdaroglu, A. Karakaya, P. Jaruga, G. Slupphaug, H.E. Krokan. Novel activities of human uracil DNA N-glycosylase for cytosine-derived products of oxidative DNA damage, Nucleic Acids Res. 24 (1996) 418-422. [261] Q. An, P. Robins, T. Lindahl, D.E. Barnes. C --> T mutagenesis and gamma- radiation sensitivity due to deficiency in the Smug1 and Ung DNA glycosylases, EMBO J 24 (2005) 2205-2213.

74

[262] P. Auffret van der Kemp, D. Thomas, R. Barbey, R. De Oliveira, S. Boiteux. Cloning and expression in Escherichia coli of the OGG1 gene of Saccharomyces cerevisiae, which codes for a DNA glycosylase that excises 7,8-dihydro-8- oxoguanine and 2, 6-diamino-4-hydroxy-5-N-methylformamidopyrimidine, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 5197-5202. [263] S. Boiteux, J.P. Radicella. The human OGG1 gene: structure, functions, and its implication in the process of carcinogenesis, Arch. Biochem. Biophys. 377 (2000) 1-8. [264] C. Dherin, M. Dizdaroglu, H. Doerflinger, S. Boiteux, J.P. Radicella. Repair of oxidative DNA damage in Drosophila melanogaster: identification and characterization of dOgg1, a second DNA glycosylase activity for 8- hydroxyguanine and formamidopyrimidines, Nucleic Acids Res. 28 (2000) 4583- 4592. [265] M.V. Garcia-Ortiz, R.R. Ariza, T. Roldan-Arjona. An OGG1 orthologue encoding a functional 8-oxoguanine DNA glycosylase/lyase in Arabidopsis thaliana, Plant Mol. Biol. 47 (2001) 795-804. [266] S. Boiteux, L. Gellon, N. Guibourt. Repair of 8-oxoguanine in Saccharomyces cerevisiae: interplay of DNA repair and replication mechanisms, Free Radic. Biol. Med. 32 (2002) 1244-1253. [267] B. Karahalil, P.M. Girard, S. Boiteux, M. Dizdaroglu. Substrate specificity of the Ogg1 protein of Saccharomyces cerevisiae: excision of guanine lesions produced in DNA by ionizing radiation- or hydrogen peroxide/metal ion-generated free radicals, Nucleic Acids Res. 26 (1998) 1228-1233. [268] M. Audebert, J.P. Radicella, M. Dizdaroglu. Effect of single mutations in the OGG1 gene found in human tumors on the substrate specificity of the ogg1 protein, Nucleic Acids Res. 28 (2000) 2672-2678. [269] T. Morales-Ruiz, M. Birincioglu, P. Jaruga, H. Rodriguez, T. Roldan-Arjona, M. Dizdaroglu. Arabidopsis thaliana Ogg1 protein excises 8-hydroxyguanine and 2,6- diamino-4-hydroxy-5-formamidopyrimidine from oxidatively damaged DNA containing multiple lesions, Biochemistry 42 (2003) 3089-3095. [270] M. Takao, H. Aburatani, K. Kobayashi, A. Yasui. Mitochondrial targeting of human DNA glycosylases for repair of oxidative DNA damage, Nucleic Acids Res. 26 (1998) 2917-2922. [271] K. Nishioka, T. Ohtsubo, H. Oda, T. Fujiwara, D. Kang, K. Sugimachi, Y. Nakabeppu. Expression and differential intracellular localization of two major forms of human 8-oxoguanine DNA glycosylase encoded by alternatively spliced OGG1 mRNAs, Mol. Biol. Cell 10 (1999) 1637-1652. [272] T. Kohno, K. Shinmura, M. Tosaka, M. Tani, S.R. Kim, H. Sugimura, T. Nohmi, H. Kasai, J. Yokota. Genetic polymorphisms and alternative splicing of the hOGG1 gene, that is involved in the repair of 8-hydroxyguanine in damaged DNA, Oncogene 16 (1998) 3219-3225. [273] H. Blons, J.P. Radicella, O. Laccourreye, D. Brasnu, P. Beaune, S. Boiteux, P. Laurent-Puig. Frequent allelic loss at chromosome 3p distinct from genetic alterations of the 8-oxoguanine DNA glycosylase 1 gene in head and neck cancer, Mol. Carcinog. 26 (1999) 254-260. [274] K. Shinmura, T. Kohno, H. Kasai, K. Koda, H. Sugimura, J. Yokota. Infrequent mutations of the hOGG1 gene, that is involved in the excision of 8-hydroxyguanine in damaged DNA, in human gastric cancer, Jpn. J. Cancer Res. 89 (1998) 825-828. [275] M. Audebert, S. Chevillard, C. Levalois, G. Gyapay, A. Vieillefond, J. Klijanienko, P. Vielh, A.K. El Naggar, S. Oudard, S. Boiteux, J.P. Radicella. Alterations of the

75

DNA repair gene OGG1 in human clear cell carcinomas of the kidney, Cancer Res. 60 (2000) 4740-4744. [276] M.M. Slupska, W.M. Luther, J.H. Chiang, H. Yang, J.H. Miller. Functional expression of hMYH, a human homolog of the Escherichia coli MutY protein, J Bacteriol. 181 (1999) 6210-6213. [277] M.A. Pope, S.S. David. DNA damage recognition and repair by the murine MutY homologue, DNA Repair (Amst) 4 (2005) 91-102. [278] T.K. Hazra, T. Izumi, I. Boldogh, B. Imhoff, Y.W. Kow, P. Jaruga, M. Dizdaroglu, S. Mitra. Identification and characterization of a human DNA glycosylase for repair of modified bases in oxidatively damaged DNA, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 3523-3528. [279] V. Bandaru, S. Sunkara, S.S. Wallace, J.P. Bond. A novel human DNA glycosylase that removes oxidative DNA damage and is homologous to Escherichia coli endonuclease VIII, DNA Repair (Amst) 1 (2002) 517-529. [280] T.K. Hazra, Y.W. Kow, Z. Hatahet, B. Imhoff, I. Boldogh, S.K. Mokkapati, S. Mitra, T. Izumi. Identification and characterization of a novel human DNA glycosylase for repair of cytosine-derived lesions, J. Biol. Chem. 277 (2002) 30417- 30420. [281] I. Morland, V. Rolseth, L. Luna, T. Rognes, M. Bjoras, E. Seeberg. Human DNA glycosylases of the bacterial Fpg/MutM superfamily: an alternative pathway for the repair of 8-oxoguanine and other oxidation products in DNA, Nucleic Acids Res. 30 (2002) 4926-4936. [282] M. Takao, S. Kanno, K. Kobayashi, Q.M. Zhang, S. Yonei, G.T. van der Horst, A. Yasui. A back-up glycosylase in Nth1 knock-out mice is a functional Nei (endonuclease VIII) homologue, J. Biol. Chem. 277 (2002) 42205-42213. [283] K. Torisu, D. Tsuchimoto, Y. Ohnishi, Y. Nakabeppu. Hematopoietic tissue-specific expression of mouse Neil3 for endonuclease VIII-like protein, J. Biochem. 138 (2005) 763-772. [284] S.S. Wallace, V. Bandaru, S.D. Kathe, J.P. Bond. The enigma of endonuclease VIII, DNA Repair (Amst) 2 (2003) 441-453. [285] M. Liu, V. Bandaru, A. Holmes, A.M. Averill, W. Cannan, S.S. Wallace. Expression and purification of active mouse and human NEIL3 proteins, Protein Expr. Purif. 84 (2012) 130-139. [286] L. Wiederhold, J.B. Leppard, P. Kedar, F. Karimi-Busheri, A. Rasouli-Nia, M. Weinfeld, A.E. Tomkinson, T. Izumi, R. Prasad, S.H. Wilson, S. Mitra, T.K. Hazra. AP endonuclease-independent DNA base excision repair in human cells, Mol. Cell 15 (2004) 209-220. [287] H. Dou, C.A. Theriot, A. Das, M.L. Hegde, Y. Matsumoto, I. Boldogh, T.K. Hazra, K.K. Bhakat, S. Mitra. Interaction of the human DNA glycosylase NEIL1 with proliferating cell nuclear antigen. The potential for replication-associated repair of oxidized bases in mammalian genomes, J. Biol. Chem. 283 (2008) 3130-3140. [288] M.L. Hegde, C.A. Theriot, A. Das, P.M. Hegde, Z. Guo, R.K. Gary, T.K. Hazra, B. Shen, S. Mitra. Physical and functional interaction between human oxidized base- specific DNA glycosylase NEIL1 and flap endonuclease 1, J. Biol. Chem. 283 (2008) 27028-27037. [289] C.A. Theriot, M.L. Hegde, T.K. Hazra, S. Mitra. RPA physically interacts with the human DNA glycosylase NEIL1 to regulate excision of oxidative DNA base damage in primer-template structures, DNA Repair (Amst) 9 (2010) 643-652. [290] P. Jaruga, M. Birincioglu, T.A. Rosenquist, M. Dizdaroglu. Mouse NEIL1 protein is specific for excision of 2,6-diamino-4-hydroxy-5-formamidopyrimidine and 4,6-

76

diamino-5-formamidopyrimidine from oxidatively damaged DNA, Biochemistry 43 (2004) 15909-15914. [291] M. Muftuoglu, N.C. de Souza-Pinto, A. Dogan, M. Aamann, T. Stevnsner, I. Rybanska, G. Kirkali, M. Dizdaroglu, V.A. Bohr. Cockayne syndrome group B protein stimulates repair of formamidopyrimidines by NEIL1 DNA glycosylase, J. Biol. Chem. 284 (2009) 9270-9279. [292] L.M. Roy, P. Jaruga, T.G. Wood, A.K. McCullough, M. Dizdaroglu, R.S. Lloyd. Human polymorphic variants of the NEIL1 DNA glycosylase, J. Biol. Chem. 282 (2007) 15790-15798. [293] M. Liu, V. Bandaru, J.P. Bond, P. Jaruga, X. Zhao, P.P. Christov, C.J. Burrows, C.J. Rizzo, M. Dizdaroglu, S.S. Wallace. The mouse ortholog of NEIL3 is a functional DNA glycosylase in vitro and in vivo, Proc. Natl. Acad. Sci. U. S. A 107 (2010) 4925-4930. [294] M. Dizdaroglu. Free-radical-induced formation of an 8,5'-cyclo-2'-deoxyguanosine moiety in deoxyribonucleic acid, Biochem. J. 238 (1986) 247-254. [295] I. Kuraoka, C. Bender, A. Romieu, J. Cadet, R.D. Wood, T. Lindahl. Removal of oxygen free-radical-induced 5',8-purine cyclodeoxynucleosides from DNA by the nucleotide excision-repair pathway in human cells, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 3832-3837. [296] P.J. Brooks, D.S. Wise, D.A. Berry, J.V. Kosmoski, M.J. Smerdon, R.L. Somers, H. Mackie, A.Y. Spoonde, E.J. Ackerman, K. Coleman, R.E. Tarone, J.H. Robbins. The oxidative DNA lesion 8,5'-(S)-cyclo-2'-deoxyadenosine is repaired by the nucleotide excision repair pathway and blocks gene expression in mammalian cells, J. Biol. Chem. 275 (2000) 22355-22362. [297] P. Pande, R.S. Das, C. Sheppard, Y.W. Kow, A.K. Basu. Repair efficiency of (5'S)- 8,5'-cyclo-2'-deoxyguanosine and (5'S)-8,5'-cyclo-2'-deoxyadenosine depends on the complementary base, DNA Repair (Amst) 11 (2012) 926-931. [298] K. Kropachev, S. Ding, M.A. Terzidis, A. Masi, Z. Liu, Y. Cai, M. Kolbanovskiy, C. Chatgilialoglu, S. Broyde, N.E. Geacintov, V. Shafirovich. Structural basis for the recognition of diastereomeric 5',8-cyclo-2'-deoxypurine lesions by the human nucleotide excision repair system, Nucleic Acids Res. (2014). [299] J. Tuo, C. Chen, X. Zeng, M. Christiansen, V.A. Bohr. Functional crosstalk between hOgg1 and the helicase domain of Cockayne syndrome group B protein, DNA Repair (Amst) 1 (2002) 913-927. [300] H. Dou, S. Mitra, T.K. Hazra. Repair of oxidized bases in DNA bubble structures by human DNA glycosylases NEIL1 and NEIL2, J. Biol. Chem. 278 (2003) 49679- 49684. [301] M.K. Hailer, P.G. Slade, B.D. Martin, T.A. Rosenquist, K.D. Sugden. Recognition of the oxidized lesions spiroiminodihydantoin and guanidinohydantoin in DNA by the mammalian base excision repair glycosylases NEIL1 and NEIL2, DNA Repair (Amst) 4 (2005) 41-50. [302] M. Liu, S. Doublie, S.S. Wallace. Neil3, the final frontier for the DNA glycosylases that recognize oxidative damage, Mutat. Res. 743-744 (2013) 4-11. [303] T. Lindahl. Instability and decay of the primary structure of DNA, Nature 362 (1993) 709-715. [304] B. Demple, T. Herman, D.S. Chen. Cloning and expression of APE, the cDNA encoding the major human apurinic endonuclease: definition of a family of DNA repair enzymes, Proc. Natl. Acad. Sci. U. S. A 88 (1991) 11450-11454. [305] D.M. Wilson, III, D. Barsky. The major human abasic endonuclease: formation, consequences and repair of abasic lesions in DNA, Mutat. Res. 485 (2001) 283-307.

77

[306] A.B. Robertson, A. Klungland, T. Rognes, I. Leiros. DNA repair in mammalian cells: Base excision repair: the long and short of it, Cell Mol. Life Sci. 66 (2009) 981-993. [307] M. Häring, H. Rüdiger, B. Demple, S. Boiteux, B. Epe. Recognition of oxidized abasic sites by repair endonucleases, Nucleic Acids Res. 22 (1994) 2010-2015. [308] J.D. Levin, B. Demple. In vitro detection of endonuclease IV-specific DNA damage formed by bleomycin in vivo, Nucleic Acids Res. 24 (1996) 885-889. [309] Y.J. Xu, E.Y. Kim, B. Demple. Excision of C-4'-oxidized deoxyribose lesions from double-stranded DNA by human apurinic/apyrimidinic endonuclease (Ape1 protein) and DNA polymerase beta, J. Biol. Chem. 273 (1998) 28837-28844. [310] M.S. DeMott, E. Beyret, D. Wong, B.C. Bales, J.T. Hwang, M.M. Greenberg, B. Demple. Covalent trapping of human DNA polymerase beta by the oxidative DNA lesion 2-deoxyribonolactone, J. Biol. Chem. 277 (2002) 7637-7640. [311] B. Demple, M.S. DeMott. Dynamics and diversions in base excision DNA repair of oxidized abasic lesions, Oncogene 21 (2002) 8926-8934. [312] Y.J. Xu, M.S. DeMott, J.T. Hwang, M.M. Greenberg, B. Demple. Action of human apurinic endonuclease (Ape1) on C1'-oxidized deoxyribose damage in DNA, DNA Repair (Amst) 2 (2003) 175-185. [313] M.M. Greenberg, Y.N. Weledji, K.M. Kroeger, J. Kim. In vitro replication and repair of DNA containing a C2'-oxidized abasic site, Biochemistry 43 (2004) 15217-15222. [314] M.M. Greenberg, Y.N. Weledji, J. Kim, B.C. Bales. Repair of oxidized abasic sites by exonuclease III, endonuclease IV, and endonuclease III, Biochemistry 43 (2004) 8178-8183. [315] J.S. Sung, B. Demple. Roles of base excision repair subpathways in correcting oxidized abasic sites in DNA, FEBS J 273 (2006) 1620-1629. [316] R.S. Wong, J.T. Sczepanski, M.M. Greenberg. Excision of a lyase-resistant oxidized abasic lesion from DNA, Chem. Res. Toxicol. 23 (2010) 766-770. [317] A.C. Jacobs, C.R. Kreller, M.M. Greenberg. Long patch base excision repair compensates for DNA polymerase beta inactivation by the C4'-oxidized abasic site, Biochemistry 50 (2011) 136-143. [318] H. Fung, B. Demple. Distinct roles of Ape1 protein in the repair of DNA damage induced by ionizing radiation or bleomycin, J Biol. Chem. 286 (2011) 4968-4977. [319] M.M. Greenberg. Abasic and Oxidized Abasic Site Reactivity in DNA: Enzyme Inhibition, Cross-Linking, and Nucleosome Catalyzed Reactions, Acc. Chem. Res. 47 (2014) 646-655. [320] J.S. Sung, M.S. DeMott, B. Demple. Long-patch base excision DNA repair of 2- deoxyribonolactone prevents the formation of DNA-protein cross-links with DNA polymerase beta, J. Biol. Chem. 280 (2005) 39095-39103. [321] J.S. Sung, B. Demple. Analysis of base excision DNA repair of the oxidative lesion 2-deoxyribonolactone and the formation of DNA-protein cross-links, Methods Enzymol. 408 (2006) 48-64. [322] S. Cunniffe, P. O'Neill, M.M. Greenberg, M.E. Lomax. Reduced repair capacity of a DNA clustered damage site comprised of 8-oxo-7,8-dihydro-2'-deoxyguanosine and 2-deoxyribonolactone results in an increased mutagenic potential of these lesions, Mutat. Res. 762 (2014) 32-39. [323] M. Asao, M. Kondo, H. Suemune, S.M. Hecht. Chemistry of the bleomycin-induced alkali-labile DNA lesion, J. Am. Chem. Soc. 121 (1999) 9023-9033. [324] M. Aso, K. Usui, M. Fukuda, Y. Kakihara, T. Goromaru, H. Suemune. Photochemical generation of C4'-oxidized abasic site containing

78

oligodeoxynucleotide and its efficient modification, Org. Lett. 8 (2006) 3183- 3186. [325] K. Usui, M. Aso, M. Fukuda, H. Suemune. Photochemical generation of oligodeoxynucleotide containing a C4'-oxidized abasic site and its efficient amine modification: dependence on structure and microenvironment, J Org. Chem. 73 (2008) 241-248. [326] L. Guan, M.M. Greenberg. Irreversible inhibition of DNA polymerase beta by an oxidized abasic lesion, J. Am. Chem. Soc. 132 (2010) 5004-5005. [327] L. Guan, K. Bebenek, T.A. Kunkel, M.M. Greenberg. Inhibition of short patch and long patch base excision repair by an oxidized abasic site, Biochemistry 49 (2010) 9904-9910. [328] A.J. Stevens, L. Guan, K. Bebenek, T.A. Kunkel, M.M. Greenberg. DNA polymerase lambda inactivation by oxidized abasic sites, Biochemistry 52 (2013) 975-983. [329] S. Dutta, G. Chowdhury, K.S. Gates. Interstrand cross-links generated by abasic sites in duplex DNA, J Am. Chem. Soc. 129 (2007) 1852-1853. [330] P. Regulus, B. Duroux, P.A. Bayle, A. Favier, J. Cadet, J.L. Ravanat. Oxidation of the sugar moiety of DNA by ionizing radiation or bleomycin could induce the formation of a cluster DNA lesion, Proc. Natl. Acad. Sci. U. S. A 104 (2007) 14032- 14037. [331] J.T. Sczepanski, A.C. Jacobs, M.M. Greenberg. Self-promoted DNA interstrand cross-link formation by an abasic site, J. Am. Chem. Soc. 130 (2008) 9646-9647. [332] J.T. Sczepanski, A.C. Jacobs, A. Majumdar, M.M. Greenberg. Scope and mechanism of interstrand cross-link formation by the C4'-oxidized abasic site, J Am. Chem. Soc. 131 (2009) 11132-11139. [333] L. Guan, M.M. Greenberg. DNA interstrand cross-link formation by the 1,4- dioxobutane abasic lesion, J Am. Chem. Soc. 131 (2009) 15225-15231. [334] M. Hashimoto, M.M. Greenberg, Y.W. Kow, J.T. Hwang, R.P. Cunningham. The 2- deoxyribonolactone lesion produced in DNA by neocarzinostatin and other damaging agents forms cross-links with the base-excision repair enzyme endonuclease III, J. Am. Chem. Soc. 123 (2001) 3161-3162. [335] M.Y. Son, H.I. Jun, K.G. Lee, B. Demple, J.S. Sung. Biochemical evaluation of genotoxic biomarkers for 2-deoxyribonolactone-mediated cross-link formation with histones, J Toxicol. Environ. Health A 72 (2009) 1311-1317. [336] U.A. Bohr, C.A. Smith, D.S. Okumoto, P.C. Hanawalt. DNA repair in an active gene: Removal of pyrimidine dimers from the DHFR gene of CHO cells is much more efficient than in the genome overall, Cell 40 (1985) 359-369. [337] I. Mellon, G. Spivak, P.C. Hanawalt. Selective removal of transcription-blocking DNA damage from the transcribed strand of the mammalian DHFR gene, Cell 51 (1987) 241-249. [338] I. Mellon, P.C. Hanawalt. Induction of the Escherichia coli lactose operon selectively increases repair of its transcribed DNA strand, Nature 342 (1989) 95-98. [339] A. Sancar. Mechanisms of DNA excision repair, Science 266 (1994) 1954-1956. [340] E.C. Friedberg. How nucleotide excision repair protects against cancer, Nat. Rev. Cancer 1 (2001) 22-33. [341] E.C. Friedberg. Nucleotide excision repair of DNA: The very early history, DNA Repair (Amst) 10 (2011) 668-672. [342] J.J. Lin, A. Sancar. A new mechanism for repairing oxidative damage to DNA: (A)BC excinuclease removes AP sites and thymine glycols from DNA, Biochemistry 28 (1989) 7979-7984.

79

[343] Y.W. Kow, S.S. Wallace, B. Van Houten. UvrABC nuclease complex repairs thymine glycol, an oxidative DNA base damage, Mutat. Res. 235 (1990) 147-156. [344] J.T. Reardon, T. Bessho, H.C. Kung, P.H. Bolton, A. Sancar. In vitro repair of oxidative DNA damage by human nucleotide excision repair system: Possible explanation for neurodegeneration in Xeroderma pigmentosum patients, Proc. Natl. Acad. Sci. USA 94 (1997) 9463-9468. [345] V.P. Jasti, R.S. Das, B.A. Hilton, S. Weerasooriya, Y. Zou, A.K. Basu. (5'S)-8,5'- Cyclo-2'-deoxyguanosine Is a strong block to replication, a potent pol V-dependent mutagenic lesion, and is inefficiently repaired in Escherichia coli, Biochemistry 50 (2011) 3862-3865. [346] A. Mazouzi, A. Vigouroux, B. Aikeshev, P.J. Brooks, M.K. Saparbaev, S. Morera, A.A. Ishchenko. Insight into mechanisms of 3'-5' exonuclease activity and removal of bulky 8,5'-cyclopurine adducts by apurinic/apyrimidinic endonucleases, Proc. Natl. Acad. Sci. U. S. A 110 (2013) E3071-E3080. [347] N. Kamakura, J. Yamamoto, P.J. Brooks, S. Iwai, I. Kuraoka. Effects of 5',8- cyclodeoxyadenosine triphosphates on DNA synthesis, Chem. Res. Toxicol. 25 (2012) 2718-2724. [348] S.S. Wallace. Biological consequences of free radical-damaged DNA bases, Free Radic. Biol. Med. 33 (2002) 1-14. [349] Y. Kuchino, F. Mori, H. Kasai, H. Inoue, S. Iwai, K. Miura, E. Ohtsuka, S. Nishimura. Misreading of DNA templates containing 8-hydroxydeoxyguanosine at the modified base and at adjacent residues, Nature 327 (1987) 77-79. [350] M.L. Wood, M. Dizdaroglu, E. Gajewski, J.M. Essigmann. Mechanistic studies of ionizing radiation and oxidative mutagenesis: genetic effects of a single 8- hydroxyguanine (7-hydro-8-oxoguanine) residue inserted at a unique site in a viral genome, Biochemistry 29 (1990) 7024-7032. [351] P.T. Henderson, W.L. Neeley, J.C. Delaney, F. Gu, J.C. Niles, S.S. Hah, S.R. Tannenbaum, J.M. Essigmann. Urea lesion formation in DNA as a consequence of 7,8-dihydro-8-oxoguanine oxidation and hydrolysis provides a potent source of point mutations, Chem. Res. Toxicol. 18 (2005) 12-18. [352] S. Shibutani, M. Takeshita, A.P. Grollman. Insertion of specific bases during DNA synthesis past the oxidation-damaged base 8-oxodG, Nature 349 (1991) 431-434. [353] H.J. Einolf, F.P. Guengerich. Fidelity of nucleotide insertion at 8-oxo-7,8- dihydroguanine by mammalian DNA polymerase delta. Steady-state and pre-steady- state kinetic analysis, J. Biol. Chem. 276 (2001) 3764-3771. [354] K.C. Cheng, D.S. Cahill, H. Kasai, S. Nishimura, L.A. Loeb. 8-Hydroxyguanine, an abundant form of oxidative DNA damage, causes G-T and A-C substitutions, J. Biol. Chem. 267 (1992) 166-172. [355] A.P. Grollman, M. Moriya. Mutagenesis by 8-oxoguanine: An enemy within, Trends Genet. 9 (1993) 246-249. [356] K.E. McAuley-Hecht, G.A. Leonard, N.J. Gibson, J.B. Thomson, W.P. Watson, W.N. Hunter, T. Brown. Crystal structure of a DNA duplex containing 8- hydroxydeoxyguanine-adenine base pairs, Biochemistry 33 (1994) 10266-10270. [357] C.J. Wiederholt, M.M. Greenberg. Fapy.dG instructs Klenow exo- to misincorporate deoxyadenosine, J. Am. Chem. Soc. 124 (2002) 7278-7679. [358] M.M. Greenberg. In vitro and in vivo effects of oxidative damage to deoxyguanosine, Biochem. Soc. Trans. 32 (2004) 46-50. [359] M.A. Kalam, K. Haraguchi, S. Chandani, E.L. Loechler, M. Moriya, M.M. Greenberg, A.K. Basu. Genetic effects of oxidative DNA damages: comparative

80

mutagenesis of the imidazole ring-opened formamidopyrimidines (Fapy lesions) and 8-oxo-purines in simian kidney cells, Nucleic Acids Res. 34 (2006) 2305-2315. [360] M.M. Greenberg. The formamidopyrimidines: purine lesions formed in competition with 8-oxopurines from oxidative stress, Acc. Chem. Res. 45 (2012) 588-597. [361] J.N. Patro, C.J. Wiederholt, Y.L. Jiang, J.C. Delaney, J.M. Essigmann, M.M. Greenberg. Studies on the replication of the ring opened formamidopyrimidine, Fapy.dG in Escherichia coli, Biochemistry 46 (2007) 10202-10212. [362] M. Olivier, M. Hollstein, P. Hainaut. TP53 mutations in human cancers: origins, consequences, and clinical use, Cold Spring Harb. Perspect. Biol. 2 (2010) a001008. [363] W. Guschlbauer, A.M. Duplaa, A. Guy, R. Téoule, G.V. Fazakerley. Structure and in vitro replication of DNA templates containing 7,8-dihydro-8-oxoadenine, Nucleic Acids Res. 19 (1991) 1753-1758. [364] G.A. Leonard, A. Guy, T. Brown, R. TEoule, W.N. Hunter. Conformation of guanine-8-oxoadenine base pairs in the crystal structure of d(CGCGAATT(O8A)GCG), Biochemistry 31 (1992) 8415-8420. [365] S. Shibutani, V. Bodepudi, F. Johnson, A.P. Grollman. Translesional synthesis on DNA templates containing 8-oxo-7,8-dihydrodeoxyadenosine, Biochemistry 32 (1993) 4615-4621. [366] H. Kamiya, H. Miura, N. Murata-Kamiya, H. Ishikawa, T. Sakaguchi, H. Inoue, T. Sasaki, C. Masutani, F. Hanaoka, S. Nishimura, E. Ohtsuka. 8-hydroxyadenine (7,8- dihydro-8-oxoadenine) induces misincorporation in in vitro DNA synthesis and mutations in NIH 3T3 cells, Nucleic Acids Res. 23 (1995) 2893-2899. [367] H. Kamiya. Mutagenic potentials of damaged nucleic acids produced by reactive oxygen/nitrogen species: approaches using synthetic oligonucleotides and nucleotides: survey and summary, Nucleic Acids Res. 31 (2003) 517-531. [368] M.L. Wood, A. Esteve, M.L. Morningstar, G.M. Kuziemko, J.M. Essigmann. Genetic effects of oxidative DNA damage: comparative mutagenesis of 7,8-dihydro- 8-oxoguanine and 7,8-dihydro-8-oxoadenine in Escherichia coli, Nucleic Acids Res. 20 (1992) 6023-6032. [369] X. Tan, A.P. Grollman, S. Shibutani. Comparison of the mutagenic properties of 8- oxo-7,8-dihydro-2'- deoxyadenosine and 8-oxo-7,8-dihydro-2'-deoxyguanosine DNA lesions in mammalian cells, Carcinogenesis 20 (1999) 2287-2292. [370] M.O. Delaney, C.J. Wiederholt, M.M. Greenberg. Fapy-dA induces nucleotide misincorporation tranlesionally by a DNA polymerase., Angew. Chem. Int. Ed. Engl. 41 (2002) 771-775. [371] H. Kamiya, T. Ueda, T. Ohgi, A. Matsukage, H. Kasai. Misincorporation of dAMP opposite 2-hydroxyadenine, an oxidative form of adenine, Nucleic Acids Res. 23 (1995) 761-766. [372] H. Kamiya, H. Kasai. Effects of sequence contexts on misincorporation of nucleotides opposite 2-hydroxyadenine, FEBS Lett. 391 (1996) 113-116. [373] H. Kamiya, H. Kasai. Substitution and deletion mutations induced by 2- hydroxyadenine in Escherichia coli: effects of sequence contexts in leading and lagging strands, Nucleic Acids Res. 25 (1997) 304-311. [374] H. Kamiya, H. Kasai. Mutations induced by 2-hydroxyadenine on a shuttle vector during leading and lagging strand syntheses in mammalian cells, Biochemistry 36 (1997) 11125-11130. [375] H. Kamiya, K. Miura, H. Ishikawa, H. Inoue, S. Nishimura, E. Ohtsuka. c-Ha-ras containing 8-hydroxyguanine at codon 12 induces point mutations at the modified and adjacent positions, Cancer Res. 52 (1992) 3483-3485.

81

[376] H. Kamiya, N. Murata-Kamiya, S. Koizume, H. Inoue, S. Nishimura, E. Ohtsuka. 8- hydroxyguanine (7,8-dihydro-8-oxoguanine) in hot spots of the c-Ha-ras gene: Effects of sequence contexts on mutation spectra, Carcinogenesis 16 (1995) 883- 889. [377] M. Inoue, H. Kamiya, K. Fujikawa, Y. Ootsuyama, N. Murata-Kamiya, T. Osaki, K. Yasumoto, H. Kasai. Induction of chromosomal gene mutations in Escherichia coli by direct incorporation of oxidatively damaged nucleotides. New evaluation method for mutagenesis by damaged DNA precursors in vivo, J Biol. Chem. 273 (1998) 11069-11074. [378] K. Satou, H. Harashima, H. Kamiya. Mutagenic effects of 2-hydroxy-dATP on replication in a HeLa extract: induction of substitution and deletion mutations, Nucleic Acids Res. 31 (2003) 2570-2575. [379] R.C. Hayes, L.A. Petrullo, H.M. Huang, S.S. Wallace, J.E. LeClerc. Oxidative damage in DNA. Lack of mutagenicity by thymine glycol lesions, J. Mol. Biol. 201 (1988) 239-246. [380] A.A. Purmal, G.W. Lampman, J.P. Bond, Z. Hatahet, S.S. Wallace. Enzymatic processing of uracil glycol, a major oxidative product of DNA cytosine, J. Biol. Chem. 273 (1998) 10026-10035. [381] T. Najrana, Y. Saito, F. Uraki, K. Kubo, K. Yamamoto. Spontaneous and osmium tetroxide-induced mutagenesis in an Escherichia coli strain deficient in both endonuclease III and endonuclease VIII, Mutagenesis 15 (2000) 121-125. [382] P. Aller, M.A. Rould, M. Hogg, S.S. Wallace, S. Doublie. A structural rationale for stalling of a replicative DNA polymerase at the most common oxidative thymine lesion, thymine glycol, Proc. Natl. Acad. Sci. U. S. A 104 (2007) 814-818. [383] R.C. Hayes, J.E. LeClerc. Sequence dependence for bypass of thymine glycols in DNA by DNA polymerase I, Nucleic Acids Res. 14 (1986) 1045-1061. [384] J.M. Clark, G.P. Beardsley. Template length, sequence context, and 3'-5' exonuclease activity modulate replicative bypass of thymine glycol lesions in vitro, Biochemistry 28 (1989) 775-779. [385] A.K. Basu, E.L. Loechler, S.A. Leadon, J.M. Essigmann. Genetic effects of thymine glycol: site-specific mutagenesis and molecular modeling studies, Proc. Natl. Acad. Sci. U. S. A. 86 (1989) 7677-7681. [386] D. Wang, D.A. Kreutzer, J.M. Essigmann. Mutagenicity and repair of oxidative DNA damage: insights from studies using defined lesions, Mutat. Res. Fundam. Mol. Mech. Mutagen. 400 (1998) 99-115. [387] H. Ide, Y.W. Kow, S.S. Wallace. Thymine glycols and urea residues in M13 DNA constitute replicative blocks in vitro, Nucleic Acids Res. 13 (1985) 8035-8052. [388] P. Rouet, J.M. Essigmann. Possible role for thymine glycol in the selective inhibition of DNA synthesis on oxidized DNA templates, Cancer. Res. 45 (1985) 6113-6118. [389] E. Moran, S.S. Wallace. The role of specific DNA base damages in the X-ray- induced inactivation of bacteriophage PM2, Mutat. Res. 146 (1985) 229-241. [390] J.M. Clark, G.P. Beardsley. Thymine glycol lesions terminate chain elongation by DNA polymerase I in vitro, Nucleic Acids Res. 14 (1986) 737-749. [391] J.M. McNulty, B. Jerkovic, P.H. Bolton, A.K. Basu. Replication inhibition and miscoding properties of DNA templates containing a site-specific cis-thymine glycol or urea residue, Chem. Res. Toxicol. 11 (1998) 666-673. [392] M.M. Greenberg, T.J. Matray. Inhibition of klenow fragment (exo-) catalyzed DNA polymerization by (5R)-5,6-dihydro-5-hydroxythymidine and structural analogue 5,6-dihydro-5-methylthymidine, Biochemistry 36 (1997) 14071-14079.

82

[393] H. Ide, L.A. Petrullo, Z. Hatahet, S.S. Wallace. Processing of DNA base damage by DNA polymerases. Dihydrothymine and beta-ureidoisobutyric acid as models for instructive and noninstructive lesions, J. Biol. Chem. 266 (1991) 1469-1477. [394] J. Evans, M. Maccabee, Z. Hatahet, J. Courcelle, R. Bockrath, H. Ide, S. Wallace. Thymine ring saturation and fragmentation products: Lesion bypass, misinsertion and implications for mutagenesis, Mutat. Res. Genet. Toxicol. Testing 299 (1993) 147-156. [395] A.A. Purmal, G.W. Lampman, Y.W. Kow, S.S. Wallace. The sequence context- dependent mispairing of 5-hydroxycytosine and 5-hydroxyuridine in vitro, Ann. NY Acad. Sci. 726 (1994) 361-363. [396] D.I. Feig, L.C. Sowers, L.A. Loeb. Reverse chemical mutagenesis: identification of the mutagenic lesions resulting from reactive oxygen species-mediated damage to DNA, Proc. Natl. Acad. Sci. U. S. A 91 (1994) 6609-6613. [397] D.A. Kreutzer, J.M. Essigmann. Oxidized, deaminated cytosines are a source of C -- > T transitions in vivo, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 3578-3582. [398] A.A. Purmal, Y.W. Kow, S.S. Wallace. Major oxidative products of cytosine, 5- hydroxycytosine and 5-hydroxyuracil, exhibit sequence context-dependent mispairing in vitro, Nucleic Acids Res. 22 (1994) 72-78. [399] W. Suen, T.G. Spiro, L.C. Sowers, J.R. Fresco. Identification by UV resonance Raman spectroscopy of an imino tautomer of 5-hydroxy-2'-deoxycytidine, a powerful base analog transition mutagen with a much higher unfavored tautomer frequency than that of the natural residue 2'-deoxycytidine, Proc. Natl. Acad. Sci. U. S. A 96 (1999) 4500-4505. [400] C. Greenman, P. Stephens, R. Smith, G.L. Dalgliesh, C. Hunter, G. Bignell, H. Davies, J. Teague, A. Butler, C. Stevens, S. Edkins, S. O'Meara, I. Vastrik, E.E. Schmidt, T. Avis, S. Barthorpe, G. Bhamra, G. Buck, B. Choudhury, J. Clements, J. Cole, E. Dicks, S. Forbes, K. Gray, K. Halliday, R. Harrison, K. Hills, J. Hinton, A. Jenkinson, D. Jones, A. Menzies, T. Mironenko, J. Perry, K. Raine, D. Richardson, R. Shepherd, A. Small, C. Tofts, J. Varian, T. Webb, S. West, S. Widaa, A. Yates, D.P. Cahill, D.N. Louis, P. Goldstraw, A.G. Nicholson, F. Brasseur, L. Looijenga, B.L. Weber, Y.E. Chiew, A. DeFazio, M.F. Greaves, A.R. Green, P. Campbell, E. Birney, D.F. Easton, G. Chenevix-Trench, M.H. Tan, S.K. Khoo, B.T. Teh, S.T. Yuen, S.Y. Leung, R. Wooster, P.A. Futreal, M.R. Stratton. Patterns of somatic mutation in human cancer genomes, Nature 446 (2007) 153-158. [401] T.J. McBride, B.D. Preston, L.A. Loeb. Mutagenic spectrum resulting from DNA damage by oxygen radicals, Biochemistry 30 (1991) 207-213. [402] T.A. Kunkel, S.S. Patel, K.A. Johnson. Error-prone replication of repeated DNA sequences by T7 DNA polymerase in the absence of its processivity subunit, Proc. Natl. Acad. Sci. U. S. A 91 (1994) 6830-6834. [403] L.A. Loeb, R.J. Monnat, Jr. DNA polymerases and human disease, Nat. Rev. Genet. 9 (2008) 594-604. [404] J.C. Shen, W. Rideout, P. Jones. The rate of hydrolytic deamination of 5- methylcytosine in double-stranded DNA, Nucleic Acids Res. 22 (1994) 972-976. [405] L.A. Loeb, B.D. Preston. Mutagenesis by apurinic/apyrimidinic sites, Annu. Rev. Genet. 20 (1986) 201-230. [406] L.A. Loeb, B.D. Preston, E.T. Snow, R.M. Schaaper. Apurinic sites as common intermediates in mutagenesis, Basic Life Sci. 38 (1986) 341-347. [407] V. Faure, J.F. Constant, P. Dumy, M. Saparbaev. 2'-deoxyribonolactone lesion produces G->A transitions in Escherichia coli, Nucleic Acids Res. 32 (2004) 2937- 2946.

83

[408] K.M. Kroeger, J. Kim, M.F. Goodman, M.M. Greenberg. Effects of the C4'- oxidized abasic site on replication in Escherichia coli. An unusually large deletion is induced by a small lesion, Biochemistry 43 (2004) 13621-13627. [409] K.M. Kroeger, Y.L. Jiang, Y.W. Kow, M.F. Goodman, M.M. Greenberg. Mutagenic effects of 2-deoxyribonolactone in Escherichia coli. An abasic lesion that disobeys the A-rule, Biochemistry 43 (2004) 6723-6733. [410] K.M. Kroeger, J. Kim, M.F. Goodman, M.M. Greenberg. Replication of an oxidized abasic site in Escherichia coli by a dNTP-stabilized misalignment mechanism that reads upstream and downstream nucleotides, Biochemistry 45 (2006) 5048-5056. [411] H. Huang, M.M. Greenberg. Hydrogen bonding contributes to the selectivity of nucleotide incorporation opposite an oxidized abasic lesion, J. Am. Chem. Soc. 130 (2008) 6080-6081. [412] E. Crespan, E. Pasi, S. Imoto, U. Hubscher, M.M. Greenberg, G. Maga. Human DNA polymerase beta, but not lambda, can bypass a 2-deoxyribonolactone lesion together with proliferating cell nuclear antigen, ACS Chem. Biol. 8 (2013) 336-344. [413] R.A. Bennett, P.S. Swerdlow, L.F. Povirk. Spontaneous cleavage of bleomycin- induced abasic sites in chromatin and their mutagenicity in mammalian shuttle vectors, Biochemistry 32 (1993) 3188-3195. [414] J.T. Sczepanski, R.S. Wong, J.N. McKnight, G.D. Bowman, M.M. Greenberg. Rapid DNA-protein cross-linking and strand scission by an abasic site in a nucleosome core particle, Proc. Natl. Acad. Sci. U. S. A 107 (2010) 22475-22480. [415] C. Zhou, J.T. Sczepanski, M.M. Greenberg. Mechanistic studies on histone catalyzed cleavage of apyrimidinic/apurinic sites in nucleosome core particles, J Am. Chem. Soc. 134 (2012) 16734-16741. [416] C. Zhou, M.M. Greenberg. Histone-catalyzed cleavage of nucleosomal DNA containing 2-deoxyribonolactone, J Am. Chem. Soc. 134 (2012) 8090-8093. [417] J.T. Sczepanski, C. Zhou, M.M. Greenberg. Nucleosome core particle-catalyzed strand scission at abasic sites, Biochemistry 52 (2013) 2157-2164. [418] C. Zhou, J.T. Sczepanski, M.M. Greenberg. Histone modification via rapid cleavage of C4'-oxidized abasic sites in nucleosome core particles, J Am. Chem. Soc. 135 (2013) 5274-5277. [419] L.F. Povirk, I.H. Goldberg. Base substitution mutations induced in the cI gene of lambda phage by neocarzinostatin chromophore: correlation with depyrimidination hotspots at the sequence AGC, Nucleic Acids Res. 14 (1986) 1417-1426. [420] L.F. Povirk, I.H. Goldberg. A role of oxidative DNA sugar damage in mutagenesis by neocarzinostatin and bleomycin, Biochimie 69 (1987) 815-823. [421] L.S. Kappen, I.H. Goldberg. Identification of 2-deoxyribonolactone at the site of neocarzinostatin-induced cytosine release in the sequence d(AGC), Biochemistry 28 (1989) 1027-1032. [422] I.H. Goldberg. Mechanism of neocarzinostatin action: Role of DNA microstructure in determination of chemistry of bistranded oxidative damage, Acc. Chem. Res. 24 (1991) 191-198. [423] N. Berthet, Y. Roupioz, J.F. Constant, M. Kotera, J. Lhomme. Translesional synthesis on DNA templates containing the 2'-deoxyribonolactone lesion, Nucleic Acids Res. 29 (2001) 2725-2732. [424] S. Shibutani, M. Takeshita, A.P. Grollman. Translesional synthesis on DNA templates containing a single abasic site - A mechanistic study of the "A rule", J. Biol. Chem. 272 (1997) 13916-13922.

84

[425] J.E. Bajacan, M.M. Greenberg. DNA polymerase V kinetics support the instructive nature of an oxidized abasic lesion in Escherichia coli, Biochemistry 52 (2013) 6301-6303. [426] M.M. Greenberg, Y.N. Weledji, K.M. Kroeger, J. Kim, M.F. Goodman. In vitro effects of a C4'-oxidized abasic site on DNA polymerases, Biochemistry 43 (2004) 2656-2663. [427] I. Kuraoka, P. Robins, C. Masutani, F. Hanaoka, D. Gasparutto, J. Cadet, R.D. Wood, T. Lindahl. Oxygen free radical damage to DNA. Translesion synthesis by human DNA polymerase η and resistance to exonuclease action at cyclopurine deoxynucleoside residues, J. Biol. Chem. 276 (2001) 49283-49288. [428] C. Marietta, H. Gulam, P.J. Brooks. A single 8,5'-cyclo-2'-deoxyadenosine lesion in a TATA box prevents binding of the TATA binding protein and strongly reduces transcription in vivo, DNA Repair (Amst) 1 (2002) 967-975. [429] C. Marietta, P.J. Brooks. Transcriptional bypass of bulky DNA lesions causes new mutant RNA transcripts in human cells, EMBO Rep. 8 (2007) 388-393. [430] H. Huang, R.S. Das, A.K. Basu, M.P. Stone. Structures of (5'S)-8,5'-Cyclo-2'- deoxyguanosine Mismatched with dA or dT, Chem. Res. Toxicol. 25 (2012) 478- 490. [431] B. Yuan, J. Wang, H. Cao, R. Sun, Y. Wang. High-throughput analysis of the mutagenic and cytotoxic properties of DNA lesions by next-generation sequencing, Nucleic Acids Res. 39 (2011) 5945-5954. [432] A.L. Swanson, J. Wang, Y. Wang. Accurate and efficient bypass of 8,5'- cyclopurine-2'-deoxynucleosides by human and yeast DNA polymerase eta, Chem. Res. Toxicol. 25 (2012) 1682-1691. [433] V. Pednekar, S. Weerasooriya, V.P. Jasti, A.K. Basu. Mutagenicity and genotoxicity of (5'S)-8,5'-cyclo-2'-deoxyadenosine in Escherichia coli and replication of (5'S)- 8,5'-cyclopurine-2'-deoxynucleosides in vitro by DNA polymerase IV, exo-free Klenow fragment, and Dpo4, Chem. Res. Toxicol. 27 (2014) 200-210. [434] B.T. Karwowski. Formation of 5',8-cyclo-2'-deoxyadenosine in dA:T pairs as a model of double stranded DNA: A theoretical mechanics study, Computational and Theoretical Chemistry 997 (2012) 55-62. [435] L.A. Loeb. Mutator phenotype in cancer: origin and consequences, Semin. Cancer Biol. 20 (2010) 279-280. [436] R.A. Beckman. Genetic instability of cancer: Biological predictions and clinical implications, in: S. Madhusudan and D.M. Wilson, III (Eds.), DNA Repair and Cancer From Bench to Clinic, CRC Press, Boca Raton, 2013, pp. 63-91. [437] J. Hall, D.R. English, M. Artuso, B.K. Armstrong, M. Winter. DNA repair capacity as a risk factor for non-melanocytic skin cancer--a molecular epidemiological study, Int J Cancer 58 (1994) 179-184. [438] L. Grossman, Q. Wei. DNA repair and epidemiology of basal cell carcinoma, Clin. Chem. 41 (1995) 1854-1863. [439] L. Cheng, S.A. Eicher, Z.Z. Guo, W.K. Hong, M.R. Spitz, Q.Y. Wei. Reduced DNA repair capacity in head and neck cancer patients, Cancer Epidemiol. Biomarkers Prev. 7 (1998) 465-468. [440] M. D'errico, A. Calcagnile, I. Iavarone, F. Sera, G. Baliva, L.M. Chinni, R. Corona, P. Pasquini, E. Dogliotti. Factors that influence the DNA repair capacity of normal and skin cancer-affected individuals, Cancer Epidemiol. Biomarkers Prev. 8 (1999) 553-559.

85

[441] L. Cheng, M.R. Spitz, W.K. Hong, Q. Wei. Reduced expression levels of nucleotide excision repair genes in lung cancer: a case-control analysis, Carcinogenesis 21 (2000) 1527-1530. [442] M. Berwick, P. Vineis. Markers of DNA repair and susceptibility to cancer in humans: an epidemiologic review, J. Natl. Cancer Inst. 92 (2000) 874-897. [443] S.M. Lippman, M.R. Spitz. Lung cancer chemoprevention: an integrated approach, J. Clin. Oncol. 19 (2001) 74S-82S. [444] E.L. Goode, C.M. Ulrich, J.D. Potter. Polymorphisms in DNA repair genes and associations with cancer risk, Cancer Epidemiol. Biomarkers Prev. 11 (2002) 1513- 1530. [445] E.L. Goode, A.M. Dunning, B. Kuschel, C.S. Healey, N.E. Day, B.A. Ponder, D.F. Easton, P.P. Pharoah. Effect of germ-line genetic variation on breast cancer survival in a population-based study, Cancer Res. 62 (2002) 3052-3057. [446] L. Cheng, E.M. Sturgis, S.A. Eicher, M.R. Spitz, Q. Wei. Expression of nucleotide excision repair genes and the risk for squamous cell carcinoma of the head and neck, Cancer 94 (2002) 393-397. [447] T. Paz-Elizur, M. Krupsky, S. Blumenstein, D. Elinger, E. Schechtman, Z. Livneh. DNA repair activity for oxidative damage and risk of lung cancer, J. Natl. Cancer Inst. 95 (2003) 1312-1319. [448] J.J. Hu, M.C. Hall, L. Grossman, M. Hedayati, D.L. McCullough, K. Lohman, L.D. Case. Deficient nucleotide excision repair capacity enhances human prostate cancer risk, Cancer Res. 64 (2004) 1197-1201. [449] S. Madhusudan, M.R. Middleton. The emerging role of DNA repair proteins as predictive, prognostic and therapeutic targets in cancer, Cancer Treat. Rev. 31 (2005) 603-617. [450] D.O. Kennedy, M. Agrawal, J. Shen, M.B. Terry, F.F. Zhang, R.T. Senie, G. Motykiewicz, R.M. Santella. DNA repair capacity of lymphoblastoid cell lines from sisters discordant for breast cancer, J Natl. Cancer Inst. 97 (2005) 127-132. [451] H. Morimoto, J. Tsukada, Y. Kominato, Y. Tanaka. Reduced expression of human mismatch repair genes in adult T-cell leukemia, Am. J. Hematol. 78 (2005) 100- 107. [452] J.M. Weiss, E.L. Goode, W.C. Ladiges, C.M. Ulrich. Polymorphic variation in hOGG1 and risk of cancer: a review of the functional and epidemiologic literature, Mol. Carcinog. 42 (2005) 127-141. [453] R. Liu, L.H. Yin, Y.P. Pu. Reduced expression of human DNA repair genes in esophageal squamous-cell carcinoma in china, J. Toxicol. Environ. Health A 70 (2007) 956-963. [454] T. Helleday, E. Petermann, C. Lundin, B. Hodgson, R.A. Sharma. DNA repair pathways as targets for cancer therapy, Nat. Rev. Cancer 8 (2008) 193-204. [455] Z. Sevilya, Y. Leitner-Dagan, M. Pinchev, R. Kremer, D. Elinger, H.S. Rennert, E. Schechtman, L.S. Freedman, G. Rennert, T. Paz-Elizur, Z. Livneh. Low integrated DNA repair score and lung cancer risk, Cancer Prev. Res. (Phila) 7 (2014) 398-406. [456] S. Toyokuni, K. Okamoto, J. Yodoi, H. Hiai. Persistent oxidative stress in cancer, FEBS Lett. 358 (1995) 1-3. [457] D.C. Malins, G.K. Ostrander, R. Haimanot, P. Williams. A novel DNA lesion in neoplastic livers of feral fish: 2,6-diamino-4-hydroxy-5-formamidopyrimidine, Carcinogenesis 11 (1990) 1045-1047. [458] D.C. Malins, R. Haimanot. 4,6-Diamino-5-formamidopyrimidine, 8- hydroxyguanine and 8-hydroxyadenine in DNA from neoplastic liver of English

86

sole exposed to carcinogens, Biochem. Biophys. Res. Commun. 173 (1990) 614- 619. [459] D.C. Malins, R. Haimanot. Major alterations in the nucleotide structure of DNA in cancer of the female breast, Cancer Res. 51 (1991) 5430-5432. [460] R. Olinski, T. Zastawny, J. Budzbon, J. Skokowski, W. Zegarski, M. Dizdaroglu. DNA base modifications in chromatin of human cancerous tissues, FEBS Lett. 193 (1992) 198. [461] D.C. Malins, E.H. Holmes, N.L. Polissar, S.J. Gunselman. The etiology of breast cancer: Characteristic alterations in hydroxyl radical-induced DNA base lesions during oncogenesis with potential for evaluating incidence risk, Cancer 71 (1993) 3036-3043. [462] D.C. Malins. Identification of hydroxyl radical-induced lesions in DNA base structure: Biomarkers with a putative link to cancer development, J. Toxicol. Environ. Health 40 (1993) 247-261. [463] P. Jaruga, T.H. Zastawny, J. Skokowski, M. Dizdaroglu, R. Olinski. Oxidative DNA base damage and antioxidant enzyme activities in human lung cancer, FEBS Lett. 341 (1994) 59-64. [464] S. Senturker, B. Karahalil, M. Inal, H. Yilmaz, H. Muslumanoglu, G. Gedikoglu, M. Dizdaroglu. Oxidative DNA base damage and antioxidant enzyme levels in childhood acute lymphoblastic leukemia, FEBS Lett. 416 (1997) 286-290. [465] E. Mambo, S.G. Nyaga, V.A. Bohr, M.K. Evans. Defective repair of 8- hydroxyguanine in mitochondria of MCF-7 and MDA-MB-468 human breast cancer cell lines, Cancer Res. 62 (2002) 1349-1355. [466] D.C. Malins, P.M. Johnson, E.A. Barker, N.L. Polissar, T.M. Wheeler, K.M. Anderson. Cancer-related changes in prostate DNA as men age and early identification of metastasis in primary prostate tumors, Proc. Natl. Acad. Sci. U. S. A 100 (2003) 5401-5406. [467] A.R. Trzeciak, S.G. Nyaga, P. Jaruga, A. Lohani, M. Dizdaroglu, M.K. Evans. Cellular repair of oxidatively induced DNA base lesions is defective in prostate cancer cell lines, PC-3 and DU-145, Carcinogenesis 25 (2004) 1-12. [468] E. Mambo, A. Chatterjee, N.C. de Souza-Pinto, S. Mayard, B.A. Hogue, M.O. Hoque, M. Dizdaroglu, V.A. Bohr, D. Sidransky. Oxidized guanine lesions and hOgg1 activity in lung cancer, Oncogene 24 (2005) 4496-4508. [469] D.C. Malins, K.M. Anderson, P. Jaruga, C.R. Ramsey, N.K. Gilman, V.M. Green, S.W. Rostad, J.T. Emerman, M. Dizdaroglu. Oxidative changes in the DNA of stroma and epithelium from the female breast: Potential implications for breast cancer, Cell Cycle 5 (2006) 1629-1632. [470] S.G. Nyaga, A. Lohani, P. Jaruga, A.R. Trzeciak, M. Dizdaroglu, M.K. Evans. Reduced repair of 8-hydroxyguanine in the human breast cancer cell line, HCC1937, BMC. Cancer 6 (2006) 297. [471] S. Madhusudan, I.D. Hickson. DNA repair inhibition: a selective tumour targeting strategy, Trends Mol. Med. 11 (2005) 503-511. [472] P. Vodicka, R. Stetina, V. Polakova, E. Tulupova, A. Naccarati, L. Vodickova, R. Kumar, M. Hanova, B. Pardini, J. Slyskova, L. Musak, P.G. De, P. Soucek, K. Hemminki. Association of DNA repair polymorphisms with DNA repair functional outcomes in healthy human subjects, Carcinogenesis 28 (2007) 657-664. [473] T. Paz-Elizur, Z. Sevilya, Y. Leitner-Dagan, D. Elinger, L.C. Roisman, Z. Livneh. DNA repair of oxidative DNA damage in human carcinogenesis: potential application for cancer risk assessment and prevention, Cancer Lett. 266 (2008) 60- 72.

87

[474] M. D'errico, E. Parlanti, E. Dogliotti. Mechanism of oxidative DNA damage repair and relevance to human pathology, Mutat. Res. 659 (2008) 4-14. [475] A.A. Nemec, S.S. Wallace, J.B. Sweasy. Variant base excision repair proteins: contributors to genomic instability, Semin. Cancer Biol. 20 (2010) 320-328. [476] D.M. Wilson, III, D. Kim, B.R. Berquist, A.J. Sigurdson. Variation in base excision repair capacity, Mutat. Res. 711 (2011) 100-112. [477] S.S. Wallace, D.L. Murphy, J.B. Sweasy. Base excision repair and cancer, Cancer Lett. 327 (2012) 73-89. [478] R. Rosell, R.V. Lord, M. Taron, N. Reguart. DNA repair and cisplatin resistance in non-small-cell lung cancer, Lung Cancer 38 (2002) 217-227. [479] T. Helleday. Amplifying tumour-specific replication lesions by DNA repair inhibitors - a new era in targeted cancer therapy, Eur. J. Cancer 44 (2008) 921-927. [480] C. Perry, R. Sultana, S. Madhusudan. Personalized cancer medicine: DNA repair alterations are promising predictive biomarkers in cancer, in: M.R. Kelley (Ed.), DNA Repair in Cancer Therapy: Molecular Targets and Clinical Applications, Elsevier, Amsterdam, 2012, pp. 257-282. [481] K. Schmid, J. Nair, G. Winde, I. Velic, H. Bartsch. Increased levels of promutagenic etheno-DNA adducts in colonic polyps of FAP patients, Int. J. Cancer 87 (2000) 1- 4. [482] T. Obtulowicz, M. Swoboda, E. Speina, D. Gackowski, R. Rozalski, A. Siomek, J. Janik, B. Janowska, J.M. Ciesla, A. Jawien, Z. Banaszkiewicz, J. Guz, T. Dziaman, A. Szpila, R. Olinski, B. Tudek. Oxidative stress and 8-oxoguanine repair are enhanced in colon adenoma and carcinoma patients, Mutagenesis 25 (2010) 463- 471. [483] S.S. Wallace. Base excision repair: A critical player in many games, DNA Repair (Amst) (2014). [484] S. Chevillard, J.P. Radicella, C. Levalois, J. Lebeau, M.F. Poupon, S. Oudard, B. Dutrillaux, S. Boiteux. Mutations in OGG1, a gene involved in the repair of oxidative DNA damage, are found in human lung and kidney tumours, Oncogene 16 (1998) 3083-3086. [485] M. Pieretti, N.H. Khattar, S.A. Smith. Common polymorphisms and somatic mutations in human base excision repair genes in ovarian and endometrial cancers, Mutat. Res. 432 (2001) 53-59. [486] G. Mao, X. Pan, B.B. Zhu, Y. Zhang, F. Yuan, J. Huang, M.A. Lovell, M.P. Lee, W.R. Markesbery, G.M. Li, L. Gu. Identification and characterization of OGG1 mutations in patients with Alzheimer's disease, Nucleic Acids Res. 35 (2007) 2759- 2766. [487] R. Lu, H.M. Nash, G.L. Verdine. A mammalian DNA repair enzyme that excises oxidatively damaged guanines maps to a locus frequently lost in lung cancer, Curr. Biol. 7 (1997) 397-407. [488] T. Ishida, R. Takashima, M. Fukayama, C. Hamada, Y. Hippo, T. Fujii, S. Moriyama, C. Matsuba, Y. Nakahori, H. Morita, Y. Yazaki, T. Kodama, S. Nishimura, H. Aburatani. New DNA polymorphisms of human MMH/OGG1 gene: prevalence of one polymorphism among lung-adenocarcinoma patients in Japanese, Int. J. Cancer 80 (1999) 18-21. [489] H. Sugimura, T. Kohno, K. Wakai, K. Nagura, K. Genka, H. Igarashi, B.J. Morris, S. Baba, Y. Ohno, C. Gao, Z. Li, J. Wang, T. Takezaki, K. Tajima, T. Varga, T. Sawaguchi, J.K. Lum, J.J. Martinson, S. Tsugane, T. Iwamasa, K. Shinmura, J. Yokota. hOGG1 Ser326Cys polymorphism and lung cancer susceptibility, Cancer Epidemiol. Biomarkers Prev. 8 (1999) 669-674.

88

[490] H. Wikman, A. Risch, F. Klimek, P. Schmezer, B. Spiegelhalder, H. Dienemann, K. Kayser, V. Schulz, P. Drings, H. Bartsch. hOGG1 polymorphism and loss of heterozygosity (LOH): significance for lung cancer susceptibility in a caucasian population, Int. J. Cancer 88 (2000) 932-937. [491] D.Y. Xing, W. Tan, N. Song, D.X. Lin. Ser326Cys polymorphism in hOGG1 gene and risk of esophageal cancer in a Chinese population, Int. J. Cancer 95 (2001) 140- 143. [492] Y.J. Park, E.Y. Choi, J.Y. Choi, J.G. Park, H.J. You, M.H. Chung. Genetic changes of hOGG1 and the activity of oh8Gua glycosylase in colon cancer, Eur. J. Cancer 37 (2001) 340-346. [493] D.Y. Xing, W. Tan, N. Song, D.X. Lin. Ser326Cys polymorphism in hOGG1 gene and risk of esophageal cancer in a Chinese population, Int J Cancer 95 (2001) 140- 143. [494] A. Elahi, Z. Zheng, J. Park, K. Eyring, T. McCaffrey, P. Lazarus. The human OGG1 DNA repair enzyme and its association with orolaryngeal cancer risk, Carcinogenesis 23 (2002) 1229-1234. [495] H. Ito, N. Hamajima, T. Takezaki, K. Matsuo, K. Tajima, S. Hatooka, T. Mitsudomi, M. Suyama, S. Sato, R. Ueda. A limited association of OGG1 Ser326Cys polymorphism for adenocarcinoma of the lung, J. Epidemiol. 12 (2002) 258-265. [496] M.L. Le, T. Donlon, A. Lum-Jones, A. Seifried, L.R. Wilkens. Association of the hOGG1 Ser326Cys polymorphism with lung cancer risk, Cancer Epidemiol. Biomarkers Prev. 11 (2002) 409-412. [497] M. Audebert, J.B. Charbonnier, S. Boiteux, J.P. Radicella. Mitochondrial targeting of human 8-oxoguanine DNA glycosylase hOGG1 is impaired by a somatic mutation found in kidney cancer, DNA Repair (Amst) 1 (2002) 497-505. [498] H. Tsukino, T. Hanaoka, T. Otani, M. Iwasaki, M. Kobayashi, M. Hara, S. Natsukawa, K. Shaura, Y. Koizumi, Y. Kasuga, S. Tsugane. hOGG1 Ser326Cys polymorphism, interaction with environmental exposures, and gastric cancer risk in Japanese populations, Cancer Sci. 95 (2004) 977-983. [499] R.J. Hung, J. Hall, P. Brennan, P. Boffetta. Genetic polymorphisms in the base excision repair pathway and cancer risk: a HuGE review, Am. J. Epidemiol. 162 (2005) 925-942. [500] R. Hansen, M. Saebo, C.F. Skjelbred, B.A. Nexo, P.C. Hagen, G. Bock, L. Bowitz, I, E. Johnson, S. Aase, I.L. Hansteen, U. Vogel, E.H. Kure. GPX Pro198Leu and OGG1 Ser326Cys polymorphisms and risk of development of colorectal adenomas and colorectal cancer, Cancer Lett. 229 (2005) 85-91. [501] Y. Niwa, K. Matsuo, H. Ito, K. Hirose, K. Tajima, T. Nakanishi, A. Nawa, K. Kuzuya, A. Tamakoshi, N. Hamajima. Association of XRCC1 Arg399Gln and OGG1 Ser326Cys polymorphisms with the risk of cervical cancer in Japanese subjects, Gynecol. Oncol. 99 (2005) 43-49. [502] T. Poplawski, M. Arabski, D. Kozirowska, M. Blasinska-Morawiec, Z. Morawiec, A. Morawiec-Bajda, G. Klupinska, A. Jeziorski, J. Chojnacki, J. Blasiak. DNA damage and repair in gastric cancer--a correlation with the hOGG1 and RAD51 genes polymorphisms, Mutat. Res. 601 (2006) 83-91. [503] T. Kohno, H. Kunitoh, K. Toyama, S. Yamamoto, A. Kuchiba, D. Saito, N. Yanagitani, S. Ishihara, R. Saito, J. Yokota. Association of the OGG1-Ser326Cys polymorphism with lung adenocarcinoma risk, Cancer Sci. 97 (2006) 724-728.

89

[504] X. Jiao, J. Huang, S. Wu, M. Lv, Y. Hu, Jianfu, X. Su, C. Luo, B. Ce. hOGG1 Ser326Cys polymorphism and susceptibility to gallbladder cancer in a Chinese population, Int. J. Cancer 121 (2007) 501-505. [505] K. Arizono, Y. Osada, Y. Kuroda. DNA repair gene hOGG1 codon 326 and XRCC1 codon 399 polymorphisms and bladder cancer risk in a Japanese population, Jpn. J. Clin. Oncol. 38 (2008) 186-191. [506] L. Hatt, S. Loft, L. Risom, P. Moller, M. Sorensen, O. Raaschou-Nielsen, K. Overvad, A. Tjonneland, U. Vogel. OGG1 expression and OGG1 Ser326Cys polymorphism and risk of lung cancer in a prospective study, Mutat. Res. 639 (2008) 45-54. [507] F. Farinati, R. Cardin, M. Bortolami, D. Nitti, D. Basso, B.M. de, M. Cassaro, A. Sergio, M. Rugge. Oxidative DNA damage in gastric cancer: CagA status and OGG1 gene polymorphism, Int. J. Cancer 123 (2008) 51-55. [508] T. Okasaka, K. Matsuo, T. Suzuki, H. Ito, S. Hosono, T. Kawase, M. Watanabe, Y. Yatabe, T. Hida, T. Mitsudomi, H. Tanaka, K. Yokoi, K. Tajima. hOGG1 Ser326Cys polymorphism and risk of lung cancer by histological type, J Hum. Genet. 54 (2009) 739-745. [509] C.J. Liu, T.C. Hsia, R.Y. Tsai, S.S. Sun, C.H. Wang, C.C. Lin, C.W. Tsai, C.Y. Huang, C.M. Hsu, D.T. Bau. The joint effect of hOGG1 single nucleotide polymorphism and smoking habit on lung cancer in Taiwan, Anticancer Res. 30 (2010) 4141-4145. [510] M. Stanczyk, T. Sliwinski, M. Cuchra, M. Zubowska, A. Bielecka-Kowalska, M. Kowalski, J. Szemraj, W. Mlynarski, I. Majsterek. The association of polymorphisms in DNA base excision repair genes XRCC1, OGG1 and MUTYH with the risk of childhood acute lymphoblastic leukemia, Mol. Biol. Rep. 38 (2011) 445-451. [511] K. Zhao, F. Qin, B. Yan, Q. Wu, M. Cao, Z. Wang, C. Zhang. hOGG1 Ser326Cys polymophism and renal cell carcinoma risk in a Chinese population, DNA Cell Biol. 30 (2011) 317-321. [512] Z. Zhang, Q. Shi, L.E. Wang, E.M. Sturgis, M.R. Spitz, A.K. El-Naggar, W.K. Hong, Q. Wei. No Association between hOGG1 Ser326Cys polymorphism and risk of squamous cell carcinoma of the head and neck, Cancer Epidemiol. Biomarkers Prev. 13 (2004) 1081-1083. [513] U. Vogel, A. Olsen, H. Wallin, K. Overvad, A. Tjonneland, B.A. Nexo. No association between OGG1 Ser326Cys and risk of basal cell carcinoma, Cancer Epidemiol. Biomarkers Prev. 13 (2004) 1680-1681. [514] K. Curtin, W.S. Samowitz, R.K. Wolff, C.M. Ulrich, B.J. Caan, J.D. Potter, M.L. Slattery. Assessing tumor mutations to gain insight into base excision repair sequence polymorphisms and smoking in colon cancer, Cancer Epidemiol. Biomarkers Prev. 18 (2009) 3384-3388. [515] J.W. Hyun, J.Y. Choi, H.H. Zeng, Y.S. Lee, H.S. Kim, S.H. Yoon, M.H. Chung. Leukemic cell line, KG-1 has a functional loss of hOGG1 enzyme due to a point mutation and 8-hydroxydeoxyguanosine can kill KG-1, Oncogene 19 (2000) 4476- 4479. [516] A.R. Dallosso, S. Dolwani, N. Jones, S. Jones, J. Colley, J. Maynard, S. Idziaszczyk, V. Humphreys, J. Arnold, A. Donaldson, D. Eccles, A. Ellis, D.G. Evans, I.M. Frayling, F.J. Hes, R.S. Houlston, E.R. Maher, M. Nielsen, S. Parry, E. Tyler, V. Moskvina, J.P. Cheadle, J.R. Sampson. Inherited predisposition to colorectal adenomas caused by multiple rare alleles of MUTYH but not OGG1, NUDT1, NTH1 or NEIL 1, 2 or 3, Gut 57 (2008) 1252-1255.

90

[517] V.S. Sidorenko, A.P. Grollman, P. Jaruga, M. Dizdaroglu, D.O. Zharkov. Substrate specificity and excision kinetics of natural polymorphic variants and phosphomimetic mutants of human 8-oxoguanine-DNA glycosylase, FEBS J. 276 (2009) 5149-5162. [518] D. Gackowski, E. Speina, M. Zielinska, J. Kowalewski, R. Rozalski, A. Siomek, T. Paciorek, B. Tudek, R. Olinski. Products of oxidative DNA damage and repair as possible biomarkers of susceptibility to lung cancer, Cancer Res. 63 (2003) 4899- 4902. [519] T. Paz-Elizur, D. Elinger, Y. Leitner-Dagan, S. Blumenstein, M. Krupsky, A. Berrebi, E. Schechtman, Z. Livneh. Development of an enzymatic DNA repair assay for molecular epidemiology studies: distribution of OGG activity in healthy individuals, DNA Repair 6 (2007) 45-60. [520] B. Tudek. Base excision repair modulation as a risk factor for human cancers, Mol. Aspects Med. 28 (2007) 258-275. [521] R.A. El-Zein, C.M. Monroy, A. Cortes, M.R. Spitz, A. Greisinger, C.J. Etzel. Rapid method for determination of DNA repair capacity in human peripheral blood lymphocytes amongst smokers, BMC Cancer 10 (2010) 439. [522] A. Klungland, I. Rosewell, S. Hollenbach, E. Larsen, G. Daly, B. Epe, E. Seeberg, T. Lindahl, D.E. Barnes. Accumulation of premutagenic DNA lesions in mice defective in removal of oxidative base damage, Proc. Natl. Acad. Sci. U. S. A 96 (1999) 13300-13305. [523] O. Minowa, T. Arai, M. Hirano, Y. Monden, S. Nakai, M. Fukuda, M. Itoh, H. Takano, Y. Hippou, H. Aburatani, K. Masumura, T. Nohmi, S. Nishimura, T. Noda. Mmh/Ogg1 gene inactivation results in accumulation of 8-hydroxyguanine in mice, Proc. Natl. Acad. Sci. U. S. A 97 (2000) 4156-4161. [524] M. Osterod, S. Hollenbach, J.G. Hengstler, D.E. Barnes, T. Lindahl, B. Epe. Age- related and tissue-specific accumulation of oxidative DNA base damage in 7,8- dihydro-8-oxoguanine-DNA glycosylase (Ogg1) deficient mice, Carcinogenesis 22 (2001) 1459-1463. [525] T. Arai, V.P. Kelly, O. Minowa, T. Noda, S. Nishimura. High accumulation of oxidative DNA damage, 8-hydroxyguanine, in Mmh/Ogg1 deficient mice by chronic oxidative stress, Carcinogenesis 23 (2002) 2005-2010. [526] T. Arai, V.P. Kelly, K. Komoro, O. Minowa, T. Noda, S. Nishimura. Cell proliferation in liver of Mmh/Ogg1-deficient mice enhances mutation frequency because of the presence of 8-hydroxyguanine in DNA, Cancer Res. 63 (2003) 4287- 4292. [527] T. Arai, V.P. Kelly, O. Minowa, T. Noda, S. Nishimura. The study using wild-type and Ogg1 knockout mice exposed to potassium bromate shows no tumor induction despite an extensive accumulation of 8-hydroxyguanine in kidney DNA, Toxicology 221 (2006) 179-186. [528] E. Larsen, K. Reite, G. Nesse, C. Gran, E. Seeberg, A. Klungland. Repair and mutagenesis at oxidized DNA lesions in the developing brain of wild-type and Ogg1-/- mice, Oncogene 25 (2006) 2425-2432. [529] K. Sakumi, Y. Tominaga, M. Furuichi, P. Xu, T. Tsuzuki, M. Sekiguchi, Y. Nakabeppu. Ogg1 knockout-associated lung tumorigenesis and its suppression by Mth1 gene disruption, Cancer Res. 63 (2003) 902-905. [530] M. Kunisada, K. Sakumi, Y. Tominaga, A. Budiyanto, M. Ueda, M. Ichihashi, Y. Nakabeppu, C. Nishigori. 8-Oxoguanine formation induced by chronic UVB exposure makes Ogg1 knockout mice susceptible to skin carcinogenesis, Cancer Res. 65 (2005) 6006-6010.

91

[531] Y. Xie, H. Yang, C. Cunanan, K. Okamoto, D. Shibata, J. Pan, D.E. Barnes, T. Lindahl, M. McIlhatton, R. Fishel, J.H. Miller. Deficiencies in mouse Myh and Ogg1 result in tumor predisposition and G to T mutations in codon 12 of the K-ras oncogene in lung tumors, Cancer Res. 64 (2004) 3096-3102. [532] M.T. Russo, L.G. De, P. Degan, E. Parlanti, E. Dogliotti, D.E. Barnes, T. Lindahl, H. Yang, J.H. Miller, M. Bignami. Accumulation of the oxidative base lesion 8- hydroxyguanine in DNA of tumor-prone mice defective in both the Myh and Ogg1 DNA glycosylases, Cancer Res. 64 (2004) 4411-4414. [533] K. Shinmura, H. Tao, M. Goto, H. Igarashi, T. Taniguchi, M. Maekawa, T. Takezaki, H. Sugimura. Inactivating mutations of the human base excision repair gene NEIL1 in gastric cancer, Carcinogenesis 25 (2004) 2311-2317. [534] M. Goto, K. Shinmura, H. Tao, S. Tsugane, H. Sugimura. Three novel NEIL1 promoter polymorphisms in gastric cancer patients, World J. Gastrointest. Oncol. 2 (2010) 117-120. [535] A.K. Maiti, I. Boldogh, H. Spratt, S. Mitra, T.K. Hazra. Mutator phenotype of mammalian cells due to deficiency of NEIL1 DNA glycosylase, an oxidized base- specific repair enzyme, DNA Repair (Amst) 7 (2008) 1213-1220. [536] T.A. Rosenquist, E. zaika, A.S. Fernandes, D.O. Zharkov, H. Miller, A.P. Grollman. The novel DNA glycosylase, NEIL1, protects mammalian cells from radiation- mediated cell death, DNA Repair (Amst) 2 (2003) 581-591. [537] A. Das, T.K. Hazra, I. Boldogh, S. Mitra, K.K. Bhakat. Induction of the human oxidized base-specific DNA glycosylase NEIL1 by reactive oxygen species, J. Biol. Chem. 280 (2005) 35272-35280. [538] M. Forsbring, E.S. Vik, B. Dalhus, T.H. Karlsen, A. Bergquist, E. Schrumpf, M. Bjoras, K.M. Boberg, I. Alseth. Catalytically impaired hMYH and NEIL1 mutant proteins identified in patients with primary sclerosing cholangitis and cholangiocarcinoma, Carcinogenesis 30 (2009) 1147-1154. [539] V. Vartanian, B. Lowell, I.G. Minko, T.G. Wood, J.D. Ceci, S. George, S.W. Ballinger, C.L. Corless, A.K. McCullough, R.S. Lloyd. The metabolic syndrome resulting from a knockout of the NEIL1 DNA glycosylase, Proc. Natl. Acad. Sci. U. S. A. 103 (2006) 1864-1869. [540] H. Mori, R. Ouchida, A. Hijikata, H. Kitamura, O. Ohara, Y. Li, X. Gao, A. Yasui, R.S. Lloyd, J.Y. Wang. Deficiency of the oxidative damage-specific DNA glycosylase NEIL1 leads to reduced germinal center B cell expansion, DNA Repair (Amst) 8 (2009) 1328-1332. [541] H. Sampath, A.K. Batra, V. Vartanian, J.R. Carmical, D. Prusak, I.B. King, B. Lowell, L.F. Earley, T.G. Wood, D.L. Marks, A.K. McCullough, R. Stephen. 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-E734. [542] R.L. Ahmed, K.H. Schmitz, K.E. Anderson, W.D. Rosamond, A.R. Folsom. The metabolic syndrome and risk of incident colorectal cancer, Cancer 107 (2006) 28- 36. [543] H.L. Lund, T.F. Wisloff, I. Holme, P. Nafstad. Metabolic syndrome predicts prostate cancer in a cohort of middle-aged Norwegian men followed for 27 years, Am. J. Epidemiol. 164 (2006) 769-774. [544] T. Sturmer, J.E. Buring, I.M. Lee, J.M. Gaziano, R.J. Glynn. Metabolic abnormalities and risk for colorectal cancer in the physicians' health study, Cancer Epidemiol. Biomarkers Prev. 15 (2006) 2391-2397.

92

[545] A. Russo, M. Autelitano, L. Bisanti. Metabolic syndrome and cancer risk, Eur. J. Cancer 44 (2008) 293-297. [546] P. Broderick, T. Bagratuni, J. Vijayakrishnan, S. Lubbe, I. Chandler, R.S. Houlston. Evaluation of NTHL1, NEIL1, NEIL2, MPG, TDG, UNG and SMUG1 genes in familial colorectal cancer predisposition, BMC. Cancer 6 (2006) 243. [547] M.T. Ocampo, W. Chaung, D.R. Marenstein, M.K. Chan, A. Altamirano, A.K. Basu, R.J. Boorstein, R.P. Cunningham, G.W. Teebor. Targeted deletion of mNth1 reveals a novel DNA repair enzyme activity, Mol. Cell Biol. 22 (2002) 6111-6121. [548] M. Takao, S. Kanno, T. Shiromoto, R. Hasegawa, H. Ide, S. Ikeda, A.H. Sarker, S. Seki, J.Z. Xing, X.C. Le, M. Weinfeld, K. Kobayashi, J. Miyazaki, M. Muijtjens, J.H. Hoeijmakers, H.G. van der, A. Yasui, A.H. Sarker. Novel nuclear and mitochondrial glycosylases revealed by disruption of the mouse Nth1 gene encoding an endonuclease III homolog for repair of thymine glycols, EMBO J. 21 (2002) 3486-3493. [549] B. Karahalil, N.C. Souza-Pinto, J.L. Parsons, R.H. Elder, V.A. Bohr. Compromised incision of oxidized pyrimidines in liver mitochondria of mice deficient in NTH1 and OGG1 glycosylases, J. Biol. Chem. 278 (2003) 33701-33707. [550] M. Goto, K. Shinmura, H. Igarashi, M. Kobayashi, H. Konno, H. Yamada, M. Iwaizumi, S. Kageyama, T. Tsuneyoshi, S. Tsugane, H. Sugimura. Altered expression of the human base excision repair gene NTH1 in gastric cancer, Carcinogenesis 30 (2009) 1345-1352. [551] S. Koketsu, T. Watanabe, H. Nagawa. Expression of DNA repair protein: MYH, NTH1, and MTH1 in colorectal cancer, Hepatogastroenterology 51 (2004) 638-642. [552] S. Xanthoudakis, T. Curran. Identification and characterization of Ref-1, a nuclear protein that facilitates AP-1 DNA-binding activity, EMBO J. 11 (1992) 653-665. [553] B. Demple, L. Harrison. Repair of oxidative damage to DNA: enzymology and biology, Annu. Rev. Biochem. 63 (1994) 915-948. [554] R. Abbotts, S. Madhusudan. Human AP endonuclease 1 (APE1): from mechanistic insights to druggable target in cancer, Cancer Treat. Rev. 36 (2010) 425-435. [555] M.L. Fishel, C. Vascotto, M.R. Kelley. DNA base excision therapeutics: Summary of targets with focus on APE1, in: S. Madhusudan and D.M. Wilson, III (Eds.), DNA Repair and Cancer From Bench to Clinic, CRC Press, Baco Raton, 2013, pp. 233-287. [556] S. Xanthoudakis, G.G. Miao, T. Curran. The redox and DNA-repair activities of Ref-1 are encoded by nonoverlapping domains, Proc. Natl. Acad. Sci. U. S. A 91 (1994) 23-27. [557] L. Gros, A.A. Ishchenko, H. Ide, R.H. Elder, M.K. Saparbaev. The major human AP endonuclease (Ape1) is involved in the nucleotide incision repair pathway, Nucleic Acids Res. 32 (2004) 73-81. [558] G. Tell, F. Quadrifoglio, C. Tiribelli, M.R. Kelley. The many functions of APE1/Ref-1: not only a DNA repair enzyme, Antioxid. Redox. Signal. 11 (2009) 601-620. [559] T. Barnes, W.C. Kim, A.K. Mantha, S.E. Kim, T. Izumi, S. Mitra, C.H. Lee. Identification of Apurinic/apyrimidinic endonuclease 1 (APE1) as the endoribonuclease that cleaves c-myc mRNA, Nucleic Acids Res. 37 (2009) 3946- 3958. [560] S. Xanthoudakis, R.J. Smeyne, J.D. Wallace, T. Curran. 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.

93

[561] L.B. Meira, S. Devaraj, G.E. Kisby, D.K. Burns, R.L. Daniel, R.E. Hammer, S. Grundy, I. Jialal, E.C. Friedberg. Heterozygosity for the mouse Apex gene results in phenotypes associated with oxidative stress, Cancer Res. 61 (2001) 5552-5557. [562] J. Huamani, C.A. McMahan, D.C. Herbert, R. Reddick, J.R. McCarrey, M.I. MacInnes, D.J. Chen, C.A. Walter. Spontaneous mutagenesis is enhanced in Apex heterozygous mice, Mol. Cell Biol. 24 (2004) 8145-8153. [563] R. Hakem. DNA-damage repair; the good, the bad, and the ugly, EMBO J 27 (2008) 589-605. [564] D.C. Cabelof. Haploinsufficiency in mouse models of DNA repair deficiency: modifiers of penetrance, Cell Mol. Life Sci. 69 (2012) 727-740. [565] H. Fung, B. Demple. A vital role for Ape1/Ref1 protein in repairing spontaneous DNA damage in human cells, Mol. Cell 17 (2005) 463-470. [566] T. Izumi, D.B. Brown, C.V. Naidu, K.K. Bhakat, M.A. Macinnes, H. Saito, D.J. Chen, S. Mitra. Two essential but distinct functions of the mammalian abasic endonuclease, Proc. Natl. Acad. Sci. U. S. A 102 (2005) 5739-5743. [567] L.J. Walker, R.B. Craig, A.L. Harris, I.D. Hickson. A role for the human DNA repair enzyme HAP1 in cellular protection against DNA damaging agents and hypoxic stress, Nucleic Acids Res. 22 (1994) 4884-4889. [568] M.R. Vasko, C. Guo, M.R. Kelley. The multifunctional DNA repair/redox enzyme Ape1/Ref-1 promotes survival of neurons after oxidative stress, DNA Repair (Amst) 4 (2005) 367-379. [569] Y. Jiang, C. Guo, M.L. Fishel, Z.Y. Wang, M.R. Vasko, M.R. Kelley. Role of APE1 in differentiated neuroblastoma SH-SY5Y cells in response to oxidative stress: use of APE1 small molecule inhibitors to delineate APE1 functions, DNA Repair (Amst) 8 (2009) 1273-1282. [570] R.A. Stetler, Y. Gao, R.S. Zukin, P.S. Vosler, L. Zhang, F. Zhang, G. Cao, M.V. Bennett, J. Chen. Apurinic/apyrimidinic endonuclease APE1 is required for PACAP-induced neuroprotection against global cerebral ischemia, Proc. Natl. Acad. Sci. U. S. A 107 (2010) 3204-3209. [571] R.M. Abbotts, P. Perry, S. Madhusudan. Human apurinic/apyrimidinic endonuclease is a novel drug target in cancer, in: S. Vengrova (Ed.), DNA Repair and Human Health, InTech, Rijeka, 2011, pp. 495-520. [572] J.L. Illuzzi, D.M. Wilson, III. Base excision repair: contribution to tumorigenesis and target in anticancer treatment paradigms, Curr. Med. Chem. 19 (2012) 3922- 3936. [573] M.Z. Hadi, M.A. Coleman, K. Fidelis, H.W. Mohrenweiser, D.M. Wilson, III. Functional characterization of Ape1 variants identified in the human population, Nucleic Acids Res. 28 (2000) 3871-3879. [574] H.W. Mohrenweiser, T. Xi, J. Vazquez-Matias, I.M. Jones. Identification of 127 amino acid substitution variants in screening 37 DNA repair genes in humans, Cancer Epidemiol. Biomarkers Prev. 11 (2002) 1054-1064. [575] J.L. Illuzzi, N.A. Harris, B.A. Manvilla, D. Kim, M. Li, A.C. Drohat, D.M. Wilson, III. Functional assessment of population and tumor-associated APE1 protein variants, PLoS. ONE. 8 (2013) e65922. [576] C. Li, Z. Liu, L.E. Wang, S.S. Strom, J.E. Lee, J.E. Gershenwald, M.I. Ross, P.F. Mansfield, J.N. Cormier, V.G. Prieto, M. Duvic, E.A. Grimm, Q. Wei. Genetic variants of the ADPRT, XRCC1 and APE1 genes and risk of cutaneous melanoma, Carcinogenesis 27 (2006) 1894-1901.

94

[577] T. Farkasova, S. Gurska, V. Witkovsky, A. Gabelova. Significance of amino acid substitution variants of DNA repair genes in radiosusceptibility of cervical cancer patients; a pilot study, Neoplasma 55 (2008) 330-337. [578] D. Gu, M. Wang, M. Wang, Z. Zhang, J. Chen. The DNA repair gene APE1 T1349G polymorphism and cancer risk: a meta-analysis of 27 case-control studies, Mutagenesis 24 (2009) 507-512. [579] B. Agachan, O. Kucukhuseyin, P. Aksoy, A. Turna, I. Yaylim, U. Gormus, A. Ergen, U. Zeybek, B. Dalan, T. Isbir. Apurinic/apyrimidinic endonuclease (APE1) gene polymorphisms and lung cancer risk in relation to tobacco smoking, Anticancer Res. 29 (2009) 2417-2420. [580] J.J. Hu, T.R. Smith, M.S. Miller, D. Lohman. Genetic regulation of ionizing radiation sensitivity and breast cancer risk, Environ. Mol. Mutagen. 39 (2002) 208- 215. [581] K. Jelonek, A. Gdowicz-Klosok, M. Pietrowska, M. Borkowska, J. Korfanty, J. Rzeszowska-Wolny, P. Widlak. Association between single-nucleotide polymorphisms of selected genes involved in the response to DNA damage and risk of colon, head and neck, and breast cancers in a Polish population, J Appl. Genet. 51 (2010) 343-352. [582] E. Canbay, B. Cakmakoglu, U. Zeybek, S. Sozen, C. Cacina, M. Gulluoglu, E. Balik, T. Bulut, S. Yamaner, D. Bugra. Association of APE1 and hOGG1 polymorphisms with colorectal cancer risk in a Turkish population, Curr. Med. Res. Opin. 27 (2011) 1295-1302. [583] S.H. Zhang, L.A. Wang, Z. Li, Y. Peng, y.p. Cun, n. Dai, y. cheng, H. Xiao, y.l. xiong, D. Wang. APE1 polymorphisms are associated with colorectal cancer susceptibility in Chinese Hans, World J. Gastroenterol. 20 (2014) 8700-8708. [584] C. Hayward, S. Colville, R.J. Swingler, D.J. Brock. Molecular genetic analysis of the APEX nuclease gene in amyotrophic lateral sclerosis, Neurology 52 (1999) 1899-1901. [585] J.J. Hu, T.R. Smith, M.S. Miller, H.W. Mohrenweiser, A. Golden, L.D. Case. Amino acid substitution variants of APE1 and XRCC1 genes associated with ionizing radiation sensitivity, Carcinogenesis 22 (2001) 917-922. [586] Y. Zhang, P.A. Newcomb, K.M. Egan, L. Titus-Ernstoff, S. Chanock, R. Welch, L.A. Brinton, J. Lissowska, A. Bardin-Mikolajczak, B. Peplonska, N. Szeszenia- Dabrowska, W. Zatonski, M. Garcia-Closas. Genetic polymorphisms in base- excision repair pathway genes and risk of breast cancer, Cancer Epidemiol. Biomarkers Prev. 15 (2006) 353-358. [587] D.F. Easton, K.A. Pooley, A.M. Dunning, P.D. Pharoah, D. Thompson, D.G. Ballinger, J.P. Struewing, J. Morrison, H. Field, R. Luben, N. Wareham, S. Ahmed, C.S. Healey, R. Bowman, K.B. Meyer, C.A. Haiman, L.K. Kolonel, B.E. Henderson, M.L. Le, P. Brennan, S. Sangrajrang, V. Gaborieau, F. Odefrey, C.Y. Shen, P.E. Wu, H.C. Wang, D. Eccles, D.G. Evans, J. Peto, O. Fletcher, N. Johnson, S. Seal, M.R. Stratton, N. Rahman, G. Chenevix-Trench, S.E. Bojesen, B.G. Nordestgaard, C.K. Axelsson, M. Garcia-Closas, L. Brinton, S. Chanock, J. Lissowska, B. Peplonska, H. Nevanlinna, R. Fagerholm, H. Eerola, D. Kang, K.Y. Yoo, D.Y. Noh, S.H. Ahn, D.J. Hunter, S.E. Hankinson, D.G. Cox, P. Hall, S. Wedren, J. Liu, Y.L. Low, N. Bogdanova, P. Schurmann, T. Dork, R.A. Tollenaar, C.E. Jacobi, P. Devilee, J.G. Klijn, A.J. Sigurdson, M.M. Doody, B.H. Alexander, J. Zhang, A. Cox, I.W. Brock, G. MacPherson, M.W. Reed, F.J. Couch, E.L. Goode, J.E. Olson, H. Meijers-Heijboer, A. van den Ouweland, A. Uitterlinden, F. Rivadeneira, R.L. Milne, G. Ribas, A. Gonzalez-Neira, J. Benitez, J.L. Hopper, M.

95

McCredie, M. Southey, G.G. Giles, C. Schroen, C. Justenhoven, H. Brauch, U. Hamann, Y.D. Ko, A.B. Spurdle, J. Beesley, X. Chen, A. Mannermaa, V.M. Kosma, V. Kataja, J. Hartikainen, N.E. Day, D.R. Cox, B.A. Ponder. Genome-wide association study identifies novel breast cancer susceptibility loci, Nature 447 (2007) 1087-1093. [588] S.N. Stacey, A. Manolescu, P. Sulem, S. Thorlacius, S.A. Gudjonsson, G.F. Jonsson, M. Jakobsdottir, J.T. Bergthorsson, J. Gudmundsson, K.K. Aben, L.J. Strobbe, D.W. Swinkels, K.C. van Engelenburg, B.E. Henderson, L.N. Kolonel, M.L. Le, E. Millastre, R. Andres, B. Saez, J. Lambea, J. Godino, E. Polo, A. Tres, S. Picelli, J. Rantala, S. Margolin, T. Jonsson, H. Sigurdsson, T. Jonsdottir, J. Hrafnkelsson, J. Johannsson, T. Sveinsson, G. Myrdal, H.N. Grimsson, S.G. Sveinsdottir, K. Alexiusdottir, J. Saemundsdottir, A. Sigurdsson, J. Kostic, L. Gudmundsson, K. Kristjansson, G. Masson, J.D. Fackenthal, C. Adebamowo, T. Ogundiran, O.I. Olopade, C.A. Haiman, A. Lindblom, J.I. Mayordomo, L.A. Kiemeney, J.R. Gulcher, T. Rafnar, U. Thorsteinsdottir, O.T. Johannsson, A. Kong, K. Stefansson. Common variants on chromosome 5p12 confer susceptibility to estrogen receptor-positive breast cancer, Nat. Genet. 40 (2008) 703-706. [589] G. Thomas, K.B. Jacobs, P. Kraft, M. Yeager, S. Wacholder, D.G. Cox, S.E. Hankinson, A. Hutchinson, Z. Wang, K. Yu, N. Chatterjee, M. Garcia-Closas, J. Gonzalez-Bosquet, L. Prokunina-Olsson, N. Orr, W.C. Willett, G.A. Colditz, R.G. Ziegler, C.D. Berg, S.S. Buys, C.A. McCarty, H.S. Feigelson, E.E. Calle, M.J. Thun, R. Diver, R. Prentice, R. Jackson, C. Kooperberg, R. Chlebowski, J. Lissowska, B. Peplonska, L.A. Brinton, A. Sigurdson, M. Doody, P. Bhatti, B.H. Alexander, J. Buring, I.M. Lee, L.J. Vatten, K. Hveem, M. Kumle, R.B. Hayes, M. Tucker, D.S. Gerhard, J.F. Fraumeni, Jr., R.N. Hoover, S.J. Chanock, D.J. Hunter. A multistage genome-wide association study in breast cancer identifies two new risk alleles at 1p11.2 and 14q24.1 (RAD51L1), Nat. Genet. 41 (2009) 579-584. [590] C. Turnbull, S. Ahmed, J. Morrison, D. Pernet, A. Renwick, M. Maranian, S. Seal, M. Ghoussaini, S. Hines, C.S. Healey, D. Hughes, M. Warren-Perry, W. Tapper, D. Eccles, D.G. Evans, M. Hooning, M. Schutte, A. van den Ouweland, R. Houlston, G. Ross, C. Langford, P.D. Pharoah, M.R. Stratton, A.M. Dunning, N. Rahman, D.F. Easton. Genome-wide association study identifies five new breast cancer susceptibility loci, Nat. Genet. 42 (2010) 504-507. [591] D.K. Braithwaite, J. Ito. Compilation, alignment, and phylogenetic relationships of DNA polymerases, Nucleic Acids Res. 21 (1993) 787-802. [592] W.A. Beard, S.H. Wilson. Structure and mechanism of DNA polymerase Beta, Chem. Rev. 106 (2006) 361-382. [593] W.A. Beard, S.H. Wilson. Structure and mechanism of DNA polymerase beta, Biochemistry 53 (2014) 2768-2780. [594] Y. Matsumoto, K. Kim. Excision of deoxyribose phosphate residues by DNA polymerase β during DNA repair, Science 269 (1995) 699-702. [595] D.K. Srivastava, B.J. Berg, R. Prasad, J.T. Molina, W.A. Beard, A.E. Tomkinson, S.H. Wilson. Mammalian abasic site base excision repair. Identification of the reaction sequence and rate-determining steps, J Biol. Chem. 273 (1998) 21203- 21209. [596] W.A. Beard, S.H. Wilson. Structural design of a eukaryotic DNA repair polymerase: DNA polymerase beta, Mutat. Res. 460 (2000) 231-244. [597] G.L. Dianov, R. Prasad, S.H. Wilson, V.A. Bohr. Role of DNA polymerase beta in the excision step of long patch mammalian base excision repair, J Biol. Chem. 274 (1999) 13741-13743.

96

[598] J.K. Horton, R. Prasad, E. Hou, S.H. Wilson. Protection against methylation- induced cytotoxicity by DNA polymerase beta-dependent long patch base excision repair, J Biol. Chem. 275 (2000) 2211-2218. [599] W.P. Osheroff, H.K. Jung, W.A. Beard, S.H. Wilson, T.A. Kunkel. The fidelity of DNA polymerase beta during distributive and processive DNA synthesis, J Biol. Chem. 274 (1999) 3642-3650. [600] W.P. Osheroff, W.A. Beard, S.H. Wilson, T.A. Kunkel. Base substitution specificity of DNA polymerase beta depends on interactions in the DNA minor groove, J Biol. Chem. 274 (1999) 20749-20752. [601] W.A. Beard, D.D. Shock, B.J. Vande Berg, S.H. Wilson. Efficiency of correct nucleotide insertion governs DNA polymerase fidelity, J Biol. Chem. 277 (2002) 47393-47398. [602] D. Starcevic, S. Dalal, J.B. Sweasy. Is there a link between DNA polymerase beta and cancer?, Cell Cycle 3 (2004) 998-1001. [603] J.B. Sweasy, T. Lang, D. Dimaio. Is base excision repair a tumor suppressor mechanism?, Cell Cycle 5 (2006) 250-259. [604] H. Gu, J.D. Marth, P.C. Orban, H. Mossmann, K. Rajewsky. Deletion of a DNA polymerase beta gene segment in T cells using cell type-specific gene targeting, Science 265 (1994) 103-106. [605] S. Banerjee, C.W. Welsch, A.R. Rao. Modulatory influence of camphor on the activities of hepatic carcinogen metabolizing enzymes and the levels of hepatic and extrahepatic reduced glutathione in mice, Cancer Lett. 88 (1995) 163-169. [606] Y. Dobashi, T. Shuin, H. Tsuruga, H. Uemura, S. Torigoe, Y. Kubota. DNA polymerase beta gene mutation in human prostate cancer, Cancer Res. 54 (1994) 2827-2829. [607] J. Matsuzaki, Y. Dobashi, H. Miyamoto, I. Ikeda, K. Fujinami, T. Shuin, Y. Kubota. DNA polymerase beta gene mutations in human bladder cancer, Mol. Carcinog. 15 (1996) 38-43. [608] H. Miyamoto, Y. Miyagi, T. Ishikawa, Y. Ichikawa, M. Hosaka, Y. Kubota. DNA polymerase beta gene mutation in human breast cancer, Int J Cancer 83 (1999) 708- 709. [609] Y. Kubota, K. Murakami-Murofushi, Y. Shimada, T. Ogiu, T. Oikawa. Reduced fidelity of DNA synthesis in cell extracts from chemically induced primary thymic lymphomas of mice, Cancer Res. 55 (1995) 3777-3780. [610] N. Bhattacharyya, H.C. Chen, S. Grundfest-Broniatowski, S. Banerjee. Alteration of hMSH2 and DNA polymerase beta genes in breast carcinomas and fibroadenomas, Biochem. Biophys. Res. Commun. 259 (1999) 429-435. [611] A. Iwanaga, M. Ouchida, K. Miyazaki, K. Hori, T. Mukai. Functional mutation of DNA polymerase beta found in human gastric cancer--inability of the base excision repair in vitro, Mutat. Res. 435 (1999) 121-128. [612] L. Wang, U. Patel, L. Ghosh, S. Banerjee. DNA polymerase beta mutations in human colorectal cancer, Cancer Res. 52 (1992) 4824-4827. [613] N. Bhattacharyya, H.C. Chen, S. Comhair, S.C. Erzurum, S. Banerjee. Variant forms of DNA polymerase beta in primary lung carcinomas, DNA Cell Biol. 18 (1999) 549-554. [614] Z. Dong, G. Zhao, Q. Zhao, H. Yang, L. Xue, X. Tan, N. Zheng. A study of DNA polymerase beta mutation in human esophageal cancer, Zhonghua Yi. Xue. Za Zhi. 82 (2002) 899-902. [615] X.H. Tan, M. Zhao, K.F. Pan, Y. Dong, B. Dong, G.J. Feng, G. Jia, Y.Y. Lu. Frequent mutation related with overexpression of DNA polymerase beta in primary

97

tumors and precancerous lesions of human stomach, Cancer Lett. 220 (2005) 101- 114. [616] R. Nowak, P. Bieganowski, R. Konopinski, J.A. Siedlecki. Alternative splicing of DNA polymerase beta mRNA is not tumor-specific, Int J Cancer 68 (1996) 199- 202. [617] T.E. Thompson, P.K. Rogan, J.I. Risinger, J.A. Taylor. Splice variants but not mutations of DNA polymerase beta are common in bladder cancer, Cancer Res. 62 (2002) 3251-3256. [618] D. Li, Y. Li, L. Jiao, D.Z. Chang, G. Beinart, R.A. Wolff, D.B. Evans, M.M. Hassan, J.L. Abbruzzese. Effects of base excision repair gene polymorphisms on pancreatic cancer survival, Int J Cancer 120 (2007) 1748-1754. [619] T. Lang, M. Maitra, D. Starcevic, S.X. Li, J.B. Sweasy. A DNA polymerase beta mutant from colon cancer cells induces mutations, Proc. Natl. Acad. Sci. U. S. A 101 (2004) 6074-6079. [620] J.B. Sweasy, T. Lang, D. Starcevic, K.W. Sun, C.C. Lai, D. Dimaio, S. Dalal. Expression of DNA polymerase {beta} cancer-associated variants in mouse cells results in cellular transformation, Proc. Natl. Acad. Sci. U. S. A 102 (2005) 14350- 14355. [621] S. Dalal, S. Hile, K.A. Eckert, K.W. Sun, D. Starcevic, J.B. Sweasy. Prostate- cancer-associated I260M variant of DNA polymerase beta is a sequence-specific mutator, Biochemistry 44 (2005) 15664-15673. [622] J. Yamtich, W.C. Speed, E. Straka, J.R. Kidd, J.B. Sweasy, K.K. Kidd. Population- specific variation in haplotype composition and heterozygosity at the POLB locus, DNA Repair (Amst) 8 (2009) 579-584. [623] Z. Guo, L. Zheng, H. Dai, M. Zhou, H. Xu, B. Shen. Human DNA polymerase beta polymorphism, Arg137Gln, impairs its polymerase activity and interaction with PCNA and the cellular base excision repair capacity, Nucleic Acids Res. 37 (2009) 3431-3441. [624] A. Matakidou, G.R. el, E.L. Webb, M.F. Rudd, H. Bridle, T. Eisen, R.S. Houlston. Genetic variation in the DNA repair genes is predictive of outcome in lung cancer, Hum. Mol. Genet. 16 (2007) 2333-2340. [625] T. Lang, S. Dalal, A. Chikova, D. Dimaio, J.B. Sweasy. The E295K DNA polymerase beta gastric cancer-associated variant interferes with base excision repair and induces cellular transformation, Mol. Cell Biol. 27 (2007) 5587-5596. [626] S. Dalal, A. Chikova, J. Jaeger, J.B. Sweasy. The Leu22Pro tumor-associated variant of DNA polymerase beta is dRP lyase deficient, Nucleic Acids Res. 36 (2008) 411-422. [627] J.M. Krahn, W.A. Beard, H. Miller, A.P. Grollman, S.H. Wilson. Structure of DNA polymerase beta with the mutagenic DNA lesion 8-oxodeoxyguanine reveals structural insights into its coding potential, Structure. 11 (2003) 121-127. [628] E. Efrati, G. Tocco, R. Eritja, S.H. Wilson, M.F. Goodman. "Action-at-a-distance" mutagenesis. 8-oxo-7, 8-dihydro-2'-deoxyguanosine causes base substitution errors at neighboring template sites when copied by DNA polymerase beta, J Biol. Chem. 274 (1999) 15920-15926. [629] H. Miller, R. Prasad, S.H. Wilson, F. Johnson, A.P. Grollman. 8-oxodGTP incorporation by DNA polymerase beta is modified by active-site residue Asn279, Biochemistry 39 (2000) 1029-1033. [630] P. Jaruga, G. Kirkali, M. Dizdaroglu. Measurement of formamidopyrimidines in DNA, Free Radic. Biol. Med. 45 (2008) 1601-1609.

98

[631] R. Cathcart, E. Schwiers, R.L. Saul, B.N. Ames. Thymine glycol and thymidine glycol in human and rat urine: a possible assay for oxidative DNA damage, Proc. Natl. Acad. Sci. USA 81 (1984) 5633-5637. [632] M.K. Shigenaga, C.J. Gimeno, B.N. Ames. Urinary 8-hydroxy-2'-deoxyguanosine as a biological marker of in vivo oxidative DNA damage, Proc. Natl. Acad. Sci. U. S. A 86 (1989) 9697-9701. [633] J.L. Ravanat, P. Guicherd, Z. Tuce, J. Cadet. Simultaneous determination of five oxidative DNA lesions in human urine, Chem. Res. Toxicol. 12 (1999) 802-808. [634] A. Weimann, D. Belling, H.E. Poulsen. Measurement of 8-oxo-2'-deoxyguanosine and 8-oxo-2'-deoxyadenosine in DNA and human urine by high performance liquid chromatography-electrospray tandem mass spectrometry, Free Radic. Biol. Med. 30 (2001) 757-764. [635] D. Gackowski, R. Rozalski, K. Roszkowski, A. Jawien, M. Foksinski, R. Olinski. 8- Oxo-7,8-dihydroguanine and 8-oxo-7,8-dihydro-2'-deoxyguanosine levels in human urine do not depend on diet, Free Radic. Res. 35 (2001) 825-832. [636] A. Weimann, D. Belling, H. Poulsen. Quantification of 8-oxo-guanine and guanine as the nucleobase, nucleoside and deoxynucleoside forms in human urine by high- performance liquid chromatography-electrospray tandem mass spectrometry, Nucleic Acids Res. 30 (2002) E7. [637] R. Rozalski, D. Gackowski, K. Roszkowski, M. Foksinski, R. Olinski. The level of 8-hydroxyguanine, a possible repair product of oxidative DNA damage, is higher in urine of cancer patients than in control subjects, Cancer Epidemiol. Biomarkers Prev. 11 (2002) 1072-1075. [638] M. Foksinski, D. Gackowski, R. Rozalski, R. Olinski. Cellular level of 8-oxo-2'- deoxyguanosine in DNA does not correlate with urinary excretion of the modified base/nucleoside, Acta Biochim. Pol. 50 (2003) 549-553. [639] A. Weimann, B. Riis, H.E. Poulsen. Oligonucleotides in human urine do not contain 8-oxo-7,8-dihydrodeoxyguanosine, Free Radic. Biol. Med. 36 (2004) 1378-1382. [640] R. Rozalski, A. Siomek, D. Gackowski, M. Foksinski, C. Gran, A. Klungland, R. Olinski. Substantial decrease of urinary 8-oxo-7,8-dihydroguanine, a product of the base excision repair pathway, in DNA glycosylase defective mice, Int. J. Biochem. Cell Biol. 37 (2005) 1331-1336. [641] de Waard H., W.J. de, J.O. Andressoo, C.T. van Oostrom, B. Riis, A. Weimann, H.E. Poulsen, S.H. van, J.H. Hoeijmakers, G.T. van der Horst. Different effects of CSA and CSB deficiency on sensitivity to oxidative DNA damage, Mol. Cell Biol. 24 (2004) 7941-7948. [642] R. Olinski, R. Rozalski, D. Gackowski, M. Foksinski, A. Siomek, M.S. Cooke. Urinary measurement of 8-OxodG, 8-OxoGua, and 5HMUra: a noninvasive assessment of oxidative damage to DNA, Antioxid. Redox. Signal. 8 (2006) 1011- 1019. [643] P. Svoboda, M. Maekawa, K. Kawai, T. Tominaga, K. Savela, H. Kasai. Urinary 8- hydroxyguanine may be a better marker of oxidative stress than 8- hydroxydeoxyguanosine in relation to the life spans of various species, Antioxid. Redox. Signal. 8 (2006) 985-992. [644] B. Malayappan, T.J. Garrett, M. Segal, C. Leeuwenburgh. Urinary analysis of 8- oxoguanine, 8-oxoguanosine, fapy-guanine and 8-oxo-2'-deoxyguanosine by high- performance liquid chromatography-electrospray tandem mass spectrometry as a measure of oxidative stress, J. Chromatogr. A 1167 (2007) 54-62. [645] M.S. Cooke, R. Olinski, S. Loft. Measurement and meaning of oxidatively modified DNA lesions in urine, Cancer Epidemiol. Biomarkers Prev. 17 (2008) 3-14.

99

[646] T. Dziaman, D. Gackowski, R. Rozalski, A. Siomek, J. Szulczynski, R. Zabielski, R. Olinski. Urinary excretion rates of 8-oxoGua and 8-oxodG and antioxidant vitamins level as a measure of oxidative status in healthy, full-term newborns, Free Radic. Res. 41 (2007) 997-1004. [647] M.S. Cooke, P.T. Henderson, M.D. Evans. Sources of extracellular, oxidatively- modified DNA lesions: implications for their measurement in urine, J. Clin. Biochem. Nutr. 45 (2009) 255-270. [648] M.D. Evans, R. Singh, V. Mistry, P.B. Farmer, M.S. Cooke. Analysis of urinary 8- oxo-7,8-dihydro-2'-deoxyguanosine by liquid chromatography-tandem mass spectrometry, Methods Mol. Biol. 610 (2010) 341-351. [649] M.D. Evans, R. Olinski, S. Loft, M.S. Cooke. Toward consensus in the analysis of urinary 8-oxo-7,8-dihydro-2'-deoxyguanosine as a noninvasive biomarker of oxidative stress, FASEB J. 24 (2010) 1249-1260. [650] L.W. Garratt, V. Mistry, R. Singh, J.K. Sandhu, B. Sheil, M.S. Cooke, P.D. Sly. Interpretation of urinary 8-oxo-7,8-dihydro-2'-deoxyguanosine is adversely affected by methodological inaccuracies when using a commercial ELISA, Free Radic. Biol. Med. (2010). [651] C.W. Hu, Y.J. Huang, Y.J. Li, M.R. Chao. Correlation between concentrations of 8- oxo-7,8-dihydro-2'-deoxyguanosine in urine, plasma and saliva measured by on-line solid-phase extraction LC-MS/MS, Clin. Chim. Acta 411 (2010) 1218-1222. [652] C.W. Hu, M.R. Chao, C.H. Sie. Urinary analysis of 8-oxo-7,8-dihydroguanine and 8-oxo-7,8-dihydro-2'-deoxyguanosine by isotope-dilution LC-MS/MS with automated solid-phase extraction: Study of 8-oxo-7,8-dihydroguanine stability, Free Radic. Biol. Med. 48 (2010) 89-97. [653] V. Mistry, F. Teichert, J.K. Sandhu, R. Singh, M.D. Evans, P.B. Farmer, M.S. Cooke. Non-invasive assessment of oxidatively damaged DNA: liquid chromatography-tandem mass spectrometry analysis of urinary 8-oxo-7,8-dihydro- 2'-deoxyguanosine, Methods Mol. Biol. 682 (2011) 279-289. [654] K. Roszkowski, R. Olinski. Urinary 8-oxoguanine as a predictor of survival in patients undergoing radiotherapy, Cancer Epidemiol. Biomarkers Prev. 21 (2012) 629-634. [655] P. Rossner, Jr., V. Mistry, R. Singh, R.J. Sram, M.S. Cooke. Urinary 8-oxo-7,8- dihydro-2'-deoxyguanosine values determined by a modified ELISA improves agreement with HPLC-MS/MS, Biochem. Biophys. Res. Commun. 440 (2013) 725- 730. [656] L. Barregard, P. Moller, T. Henriksen, V. Mistry, G. Koppen, P. Rossner, Jr., R.J. Sram, A. Weimann, H.E. Poulsen, R. Nataf, R. Andreoli, P. Manini, T. Marczylo, P. Lam, M.D. Evans, H. Kasai, K. Kawai, Y.S. Li, K. Sakai, R. Singh, F. Teichert, P.B. Farmer, R. Rozalski, D. Gackowski, A. Siomek, G.T. Saez, C. Cerda, K. Broberg, C. Lindh, M.B. Hossain, S. Haghdoost, C.W. Hu, M.R. Chao, K.Y. Wu, H. Orhan, N. Senduran, R.J. Smith, R.M. Santella, Y. Su, C. Cortez, S. Yeh, R. Olinski, S. Loft, M.S. Cooke. Human and methodological sources of variability in the measurement of urinary 8-oxo-7,8-dihydro-2'-deoxyguanosine, Antioxid. Redox. Signal. 18 (2013) 2377-2391. [657] M.S. Cooke, M.D. Evans, R. Dove, R. Rozalski, D. Gackowski, A. Siomek, J. Lunec, R. Olinski. DNA repair is responsible for the presence of oxidatively damaged DNA lesions in urine, Mutat. Res. 574 (2005) 58-66. [658] A. Siomek, J. Tujakowski, D. Gackowski, R. Rozalski, M. Foksinski, T. Dziaman, K. Roszkowski, R. Olinski. Severe oxidatively damaged DNA after cisplatin treatment of cancer patients, Int. J. Cancer 119 (2006) 2228-2230.

100

[659] S. Haghdoost, L. Sjolander, S. Czene, M. Harms-Ringdahl. The nucleotide pool is a significant target for oxidative stress, Free Radic. Biol. Med. 41 (2006) 620-626. [660] H. Hayakawa, A. Taketomi, K. Sakumi, M. Kuwano, M. Sekiguchi. Generation and elimination of 8-oxo-7,8-dihydro-2'-deoxyguanosine 5'-triphosphate, a mutagenic substrate for DNA synthesis, in human cells, Biochemistry 34 (1995) 89-95. [661] C.W. Hu, M.S. Cooke, Y.H. Tsai, M.R. Chao. 8-Oxo-7,8-dihydroguanine and 8- oxo-7,8-dihydro-2'-deoxyguanosine concentrations in various human body fluids: implications for their measurement and interpretation, Arch. Toxicol. (2014). [662] G. Damia, M. D'Incalci. Targeting DNA repair as a promising approach in cancer therapy, Eur. J Cancer 43 (2007) 1791-1801. [663] M.R. Kelley. Future directions with DNA repair inhibitors: A roadmap for disruptive approaches to cancer therapy, in: M.R. Kelley (Ed.), DNA Repair in Cancer Therapy, Molecular Targets and Clinical Applications, Elsevier, Amsterdam, 2012, pp. 301-310. [664] C. Holohan, S.S. Van, D.B. Longley, P.G. Johnston. Cancer drug resistance: an evolving paradigm, Nat. Rev. Cancer 13 (2013) 714-726. [665] B.J. Moeller, W. Arap, R. Pasqualini. Targeting synthetic lethality in DNA damage repair pathways as an anti-cancer strategy, Curr. Drug Targets. 11 (2010) 1336- 1340. [666] N. Chan, R.G. Bristow. "Contextual" synthetic lethality and/or loss of heterozygosity: tumor hypoxia and modification of DNA repair, Clin. Cancer Res. 16 (2010) 4553-4560. [667] N.J. Curtin. PARP and PARP inhibitor therapeutics, in: S. Madhusudan and D.M. Wilson, III (Eds.), DNA Repair and Cancer From Bench to Clinic, CRC Press, Baco Raton, 2013, pp. 513-563. [668] K. Aziz, S. Nowsheen, A.G. Georgakilas. Nanotechnology in cancer therapy: targeting the inhibition of key DNA repair pathways, Curr. Mol. Med. 10 (2010) 626-639. [669] B.W. Durkacz, O. Omidiji, D.A. Gray, S. Shall. (ADP-ribose)n participates in DNA excision repair, Nature 283 (1980) 593-596. [670] H. Farmer, N. McCabe, C.J. Lord, A.N. Tutt, D.A. Johnson, T.B. Richardson, M. Santarosa, K.J. Dillon, I. Hickson, C. Knights, N.M. Martin, S.P. Jackson, G.C. Smith, A. Ashworth. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy, Nature 434 (2005) 917-921. [671] H.E. Bryant, N. Schultz, H.D. Thomas, K.M. Parker, D. Flower, E. Lopez, S. Kyle, M. Meuth, N.J. Curtin, T. Helleday. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase, Nature 434 (2005) 913-917. [672] R. Plummer. Perspective on the pipeline of drugs being developed with modulation of DNA damage as a target, Clin. Cancer Res. 16 (2010) 4527-4531. [673] Y. Drew, R. Plummer. The emerging potential of poly(ADP-ribose) polymerase inhibitors in the treatment of breast cancer, Curr. Opin. Obstet. Gynecol. 22 (2010) 67-71. [674] N.J. Curtin, A. Mukhopadhyay, Y. Drew, R. Plummer. The role of PARP in DNA repair and its therapeutic exploitation, in: M.R. Kelley (Ed.), DNA Repair in Cancer Therapy, Molecular Targets and Clinical Applications, Elsevier, Amsterdam, 2012, pp. 55-73. [675] T. Helleday. Putting poly (ADP-ribose) polymerase and other DNA repair inhibitors into clinical practice, Curr. Opin. Oncol. 25 (2013) 609-614. [676] B. Tanner, S. Grimme, I. Schiffer, C. Heimerdinger, M. Schmidt, P. Dutkowski, S. Neubert, F. Oesch, A. Franzen, H. Kolbl, G. Fritz, B. Kaina, J.G. Hengstler. Nuclear

101

expression of apurinic/apyrimidinic endonuclease increases with progression of ovarian carcinomas, Gynecol. Oncol. 92 (2004) 568-577. [677] Y. Xu, D.H. Moore, J. Broshears, L. Liu, T.M. Wilson, M.R. Kelley. The apurinic/apyrimidinic endonuclease (APE/ref-1) DNA repair enzyme is elevated in premalignant and malignant cervical cancer, Anticancer Res. 17 (1997) 3713-3719. [678] S. Kakolyris, A. Giatromanolaki, M. Koukourakis, L. Kaklamanis, P. Kanavaros, I.D. Hickson, G. Barzilay, V. Georgoulias, K.C. Gatter, A.L. Harris. Nuclear localization of human AP endonuclease 1 (HAP1/Ref-1) associates with prognosis in early operable non-small cell lung cancer (NSCLC), J Pathol. 189 (1999) 351- 357. [679] F. Puglisi, G. Aprile, A.M. Minisini, F. Barbone, P. Cataldi, G. Tell, M.R. Kelley, G. Damante, C.A. Beltrami, L.C. Di. Prognostic significance of Ape1/ref-1 subcellular localization in non-small cell lung carcinomas, Anticancer Res. 21 (2001) 4041-4049. [680] M.I. Koukourakis, A. Giatromanolaki, S. Kakolyris, E. Sivridis, V. Georgoulias, G. Funtzilas, I.D. Hickson, K.C. Gatter, A.L. Harris. Nuclear expression of human apurinic/apyrimidinic endonuclease (HAP1/Ref-1) in head-and-neck cancer is associated with resistance to chemoradiotherapy and poor outcome, Int J Radiat. Oncol. Biol. Phys. 50 (2001) 27-36. [681] S. Yang, K. Irani, S.E. Heffron, F. Jurnak, F.L. Meyskens, Jr. Alterations in the expression of the apurinic/apyrimidinic endonuclease-1/redox factor-1 (APE/Ref-1) in human melanoma and identification of the therapeutic potential of resveratrol as an APE/Ref-1 inhibitor, Mol. Cancer Ther. 4 (2005) 1923-1935. [682] R.I. Al-Safi, S. Odde, Y. Shabaik, N. Neamati. Small-molecule inhibitors of APE1 DNA repair function: an overview, Curr. Mol. Pharmacol. 5 (2012) 14-35. [683] S. Madhusudan, F. Smart, P. Shrimpton, J.L. Parsons, L. Gardiner, S. Houlbrook, D.C. Talbot, T. Hammonds, P.A. Freemont, M.J. Sternberg, G.L. Dianov, I.D. Hickson. Isolation of a small molecule inhibitor of DNA base excision repair, Nucleic Acids Res. 33 (2005) 4711-4724. [684] Z. Zawahir, R. Dayam, J. Deng, C. Pereira, N. Neamati. Pharmacophore guided discovery of small-molecule human apurinic/apyrimidinic endonuclease 1 inhibitors, J Med. Chem. 52 (2009) 20-32. [685] A. Bapat, L.S. Glass, M. Luo, M.L. Fishel, E.C. Long, M.M. Georgiadis, M.R. Kelley. Novel small-molecule inhibitor of apurinic/apyrimidinic endonuclease 1 blocks proliferation and reduces viability of glioblastoma cells, J Pharmacol. Exp. Ther. 334 (2010) 988-998. [686] A. Simeonov, A. Kulkarni, D. Dorjsuren, A. Jadhav, M. Shen, D.R. McNeill, C.P. Austin, D.M. Wilson, III. Identification and characterization of inhibitors of human apurinic/apyrimidinic endonuclease APE1, PLoS. ONE. 4 (2009) e5740. [687] A. Jedinak, S. Dudhgaonkar, M.R. Kelley, D. Sliva. Apurinic/Apyrimidinic endonuclease 1 regulates inflammatory response in macrophages, Anticancer Res. 31 (2011) 379-385. [688] M.R. Kelley, M.M. Georgiadis, M.L. Fishel. APE1/Ref-1 role in redox signaling: Translational applications of targeting the redox function of the DNA repair/redox protein APE1/Ref-1, Curr. Mol. Pharmacol. 5 (2012) 36-53. [689] D.K. Srivastava, I. Husain, C.L. Arteaga, S.H. Wilson. DNA polymerase beta expression differences in selected human tumors and cell lines, Carcinogenesis 20 (1999) 1049-1054. [690] K.K. Chan, Q.M. Zhang, G.L. Dianov. Base excision repair fidelity in normal and cancer cells, Mutagenesis 21 (2006) 173-178.

102

[691] Y. Canitrot, C. Cazaux, M. Frechet, K. Bouayadi, C. Lesca, B. Salles, J.S. Hoffmann. Overexpression of DNA polymerase beta in cell results in a mutator phenotype and a decreased sensitivity to anticancer drugs, Proc. Natl. Acad. Sci. U. S. A 95 (1998) 12586-12590. [692] V. Bergoglio, M.J. Pillaire, M. Lacroix-Triki, B. Raynaud-Messina, Y. Canitrot, A. Bieth, M. Gares, M. Wright, G. Delsol, L.A. Loeb, C. Cazaux, J.S. Hoffmann. Deregulated DNA polymerase beta induces chromosome instability and tumorigenesis, Cancer Res. 62 (2002) 3511-3514. [693] N.A. Yamada, R.A. Farber. Induction of a low level of microsatellite instability by overexpression of DNA polymerase Beta, Cancer Res. 62 (2002) 6061-6064. [694] H.K. Wong, D.M. Wilson, III. XRCC1 and DNA polymerase beta interaction contributes to cellular alkylating-agent resistance and single-strand break repair, J Cell Biochem. 95 (2005) 794-804. [695] K. Yoshizawa, E. Jelezcova, A.R. Brown, J.F. Foley, A. Nyska, X. Cui, L.J. Hofseth, R.M. Maronpot, S.H. Wilson, A.R. Sepulveda, R.W. Sobol. Gastrointestinal hyperplasia with altered expression of DNA polymerase beta, PLoS. ONE. 4 (2009) e6493. [696] A.J. Fornace, Jr., B. Zmudzka, M.C. Hollander, S.H. Wilson. Induction of beta- polymerase mRNA by DNA-damaging agents in Chinese hamster ovary cells, Mol. Cell Biol. 9 (1989) 851-853. [697] D.C. Cabelof, J.J. Raffoul, S. Yanamadala, Z. Guo, A.R. Heydari. Induction of DNA polymerase beta-dependent base excision repair in response to oxidative stress in vivo, Carcinogenesis 23 (2002) 1419-1425. [698] Y.Y. Polosina, T.A. Rosenquist, A.P. Grollman, H. Miller. 'Knock down' of DNA polymerase beta by RNA interference: recapitulation of null phenotype, DNA Repair (Amst) 3 (2004) 1469-1474. [699] R.N. Trivedi, K.H. Almeida, J.L. Fornsaglio, S. Schamus, R.W. Sobol. The role of base excision repair in the sensitivity and resistance to temozolomide-mediated cell death, Cancer Res. 65 (2005) 6394-6400. [700] H.Y. Hu, J.K. Horton, M.R. Gryk, R. Prasad, J.M. Naron, D.A. Sun, S.M. Hecht, S.H. Wilson, G.P. Mullen. Identification of small molecule synthetic inhibitors of DNA polymerase beta by NMR chemical shift mapping, J Biol. Chem. 279 (2004) 39736-39744. [701] Z. Gao, D.J. Maloney, L.M. Dedkova, S.M. Hecht. Inhibitors of DNA polymerase beta: activity and mechanism, Bioorg. Med. Chem. 16 (2008) 4331-4340. [702] A.S. Jaiswal, S. Banerjee, H. Panda, C.D. Bulkin, T. Izumi, F.H. Sarkar, D.A. Ostrov, S. Narayan. A novel inhibitor of DNA polymerase beta enhances the ability of temozolomide to impair the growth of colon cancer cells, Mol. Cancer Res. 7 (2009) 1973-1983. [703] K. Barakat, J. Tuszynski. Relaxed complex scheme suggests novel inhibitors for the lyase activity of DNA polymerase beta, J Mol. Graph. Model. 29 (2011) 702-716. [704] R. Bei, L. Marzocchella, M. Turriziani. The use of temozolomide for the treatment of malignant tumors: clinical evidence and molecular mechanisms of action, Recent Pat Anticancer Drug Discov. 5 (2010) 172-187. [705] J.D. Hainsworth, T. Ervin, E. Friedman, V. Priego, P.B. Murphy, B.L. Clark, R.E. Lamar. Concurrent radiotherapy and temozolomide followed by temozolomide and sorafenib in the first-line treatment of patients with glioblastoma multiforme, Cancer 116 (2010) 3663-3669.

103

[706] A.S. Jaiswal, S. Banerjee, R. Aneja, F.H. Sarkar, D.A. Ostrov, S. Narayan. DNA polymerase beta as a novel target for chemotherapeutic intervention of colorectal cancer, PLoS. ONE. 6 (2011) e16691. [707] L.R. Hiraoka, J.J. Harrington, D.S. Gerhard, M.R. Lieber, C.L. Hsieh. Sequence of human FEN-1, a structure-specific endonuclease, and chromosomal localization of the gene (FEN1) in mouse and human, Genomics 25 (1995) 220-225. [708] A. Klungland, T. Lindahl. Second pathway for completion of human DNA base excision-repair: reconstitution with purified proteins and requirement for DNase IV (FEN1), EMBO J. 16 (1997) 3341-3348. [709] M.R. Lieber. The FEN-1 family of structure-specific in eukaryotic DNA replication, recombination and repair, BioEssays 19 (1997) 233-240. [710] B.I. Lee, D.M. Wilson, III. The RAD2 domain of human exonuclease 1 exhibits 5' to 3' exonuclease and flap structure-specific endonuclease activities, J Biol. Chem. 274 (1999) 37763-37769. [711] R.S. Williams, T.A. Kunkel. FEN nucleases: bind, bend, fray, cut, Cell 145 (2011) 171-172. [712] S.E. Tsutakawa, S. Classen, B.R. Chapados, A.S. Arvai, L.D. Finger, G. Guenther, C.G. Tomlinson, P. Thompson, A.H. Sarker, B. Shen, P.K. Cooper, J.A. Grasby, J.A. Tainer. Human flap endonuclease structures, DNA double-base flipping, and a unified understanding of the FEN1 superfamily, Cell 145 (2011) 198-211. [713] B. Shen, P. Singh, R. Liu, J. Qiu, L. Zheng, L.D. Finger, S. Alas. Multiple but dissectible functions of FEN-1 nucleases in nucleic acid processing, genome stability and diseases, BioEssays 27 (2005) 717-729. [714] G. Tell, D.M. Wilson, III, C.H. Lee. Intrusion of a DNA repair protein in the RNome world: is this the beginning of a new era?, Mol. Cell Biol. 30 (2010) 366- 371. [715] Y. Liu, H.I. Kao, R.A. Bambara. Flap endonuclease 1: a central component of DNA metabolism, Annu. Rev. Biochem. 73 (2004) 589-615. [716] E. LaTulippe, J. Satagopan, A. Smith, H. Scher, P. Scardino, V. Reuter, W.L. Gerald. Comprehensive gene expression analysis of prostate cancer reveals distinct transcriptional programs associated with metastatic disease, Cancer Res. 62 (2002) 4499-4506. [717] M. Sato, L. Girard, I. Sekine, N. Sunaga, R.D. Ramirez, C. Kamibayashi, J.D. Minna. Increased expression and no mutation of the Flap endonuclease (FEN1) gene in human lung cancer, Oncogene 22 (2003) 7243-7246. [718] S.J. Freedland, A.J. Pantuck, S.H. Paik, A. Zisman, T.G. Graeber, D. Eisenberg, W.H. McBride, D. Nguyen, C.L. Tso, A.S. Belldegrun. Heterogeneity of molecular targets on clonal cancer lines derived from a novel hormone-refractory prostate cancer tumor system, Prostate 55 (2003) 299-307. [719] A. Krause, V. Combaret, I. Iacono, B. Lacroix, C. Compagnon, C. Bergeron, S. Valsesia-Wittmann, P. Leissner, B. Mougin, A. Puisieux. Genome-wide analysis of gene expression in neuroblastomas detected by mass screening, Cancer Lett. 225 (2005) 111-120. [720] J.S. Lam, D.B. Seligson, H. Yu, A. Li, M. Eeva, A.J. Pantuck, G. Zeng, S. Horvath, A.S. Belldegrun. Flap endonuclease 1 is overexpressed in prostate cancer and is associated with a high Gleason score, BJU. Int 98 (2006) 445-451. [721] P. Singh, M. Yang, H. Dai, D. Yu, Q. Huang, W. Tan, K.H. Kernstine, D. Lin, B. Shen. Overexpression and hypomethylation of flap endonuclease 1 gene in breast and other cancers, Mol. Cancer Res. 6 (2008) 1710-1717.

104

[722] T. Nikolova, M. Christmann, B. Kaina. FEN1 is overexpressed in testis, lung and brain tumors, Anticancer Res. 29 (2009) 2453-2459. [723] L.N. Tumey, D. Bom, B. Huck, E. Gleason, J. Wang, D. Silver, K. Brunden, S. Boozer, S. Rundlett, B. Sherf, S. Murphy, T. Dent, C. Leventhal, A. Bailey, J. Harrington, Y.L. Bennani. The identification and optimization of a N-hydroxy urea series of flap endonuclease 1 inhibitors, Bioorg. Med. Chem. Lett. 15 (2005) 277- 281. [724] D. Dorjsuren, D. Kim, D.J. Maloney, D.M. Wilson, III, A. Simeonov. Complementary non-radioactive assays for investigation of human flap endonuclease 1 activity, Nucleic Acids Res. 39 (2011) e11. [725] H. Panda, A.S. Jaiswal, P.E. Corsino, M.L. Armas, B.K. Law, S. Narayan. Amino acid Asp181 of 5'-flap endonuclease 1 is a useful target for chemotherapeutic development, Biochemistry 48 (2009) 9952-9958. [726] K.J. McManus, I.J. Barrett, Y. Nouhi, P. Hieter. Specific synthetic lethal killing of RAD54B-deficient human colorectal cancer cells by FEN1 silencing, Proc. Natl. Acad. Sci. U. S. A 106 (2009) 3276-3281. [727] K.V. Huber, E. Salah, B. Radic, M. Gridling, J.M. Elkins, A. Stukalov, A.S. Jemth, C. Gokturk, K. Sanjiv, K. Stromberg, T. Pham, U.W. Berglund, J. Colinge, K.L. Bennett, J.I. Loizou, T. Helleday, S. Knapp, G. Superti-Furga. Stereospecific targeting of MTH1 by (S)-crizotinib as an anticancer strategy, Nature 508 (2014) 222-227. [728] M.T. Russo, M.F. Blasi, F. Chiera, P. Fortini, P. Degan, P. Macpherson, M. Furuichi, Y. Nakabeppu, P. Karran, G. Aquilina, M. Bignami. The oxidized deoxynucleoside triphosphate pool is a significant contributor to genetic instability in mismatch repair-deficient cells, Mol. Cell Biol. 24 (2004) 465-474. [729] K. Okamoto, S. Toyokuni, W.J. Kim, O. Ogawa, Y. Kakehi, S. Arao, H. Hiai, O. Yoshida. Overexpression of human mutT homologue gene messenger RNA in renal-cell carcinoma: evidence of persistent oxidative stress in cancer, Int J Cancer 65 (1996) 437-441. [730] C.H. Kennedy, R. Cueto, S.A. Belinsky, J.F. Lechner, W.A. Pryor. Overexpression of hMTH1 mRNA: a molecular marker of oxidative stress in lung cancer cells, FEBS Lett. 429 (1998) 17-20. [731] L. Taricani, F. Shanahan, R.H. Pierce, T.J. Guzi, D. Parry. Phenotypic enhancement of thymidylate synthetase pathway inhibitors following ablation of Neil1 DNA glycosylase/lyase, Cell Cycle 9 (2010) 4876-4883. [732] K.R. Harrap, A.L. Jackman, D.R. Newell, G.A. Taylor, L.R. Hughes, A.H. Calvert. Thymidylate synthase: a target for anticancer drug design, Adv. Enzyme Regul. 29 (1989) 161-179. [733] C.W. Carreras, D.V. Santi. The catalytic mechanism and structure of thymidylate synthase, Annu. Rev. Biochem. 64 (1995) 721-762. [734] M.G. Rose, M.P. Farrell, J.C. Schmitz. Thymidylate synthase: a critical target for cancer chemotherapy, Clin. Colorectal Cancer 1 (2002) 220-229. [735] J.J. McGuire. Anticancer : current status and future directions, Curr. Pharm. Des 9 (2003) 2593-2613. [736] S. Dubey, J.H. Schiller. Three emerging new drugs for NSCLC: , , and cetuximab, Oncologist. 10 (2005) 282-291. [737] P.M. Wilson, M.J. LaBonte, H.J. Lenz, P.C. Mack, R.D. Ladner. Inhibition of dUTPase induces synthetic lethality with thymidylate synthase-targeted therapies in non-small cell lung cancer, Mol. Cancer Ther. 11 (2012) 616-628.

105

[738] A.C. Jacobs, M.J. Calkins, A. Jadhav, D. Dorjsuren, D. Maloney, A. Simeonov, P. Jaruga, M. Dizdaroglu, A.K. McCullough, R.S. Lloyd. Inhibition of DNA glycosylases via small molecule purine analogs, PLoS. ONE. 8 (2013) e81667. [739] P.T. Reddy, P. Jaruga, B.C. Nelson, M. Lowenthal, M. Dizdaroglu. Stable isotope- labeling of DNA repair proteins, and their purification and characterization, Protein Expr. Purif. 78 (2011) 94-101. [740] M. Dizdaroglu, P.T. Reddy, P. Jaruga. Identification and quantification of DNA repair proteins by liquid chromatography/isotope-dilution tandem mass spectrometry using their fully 15N-labeled analogues as internal standards, J. Proteome. Res. 10 (2011) 3802-3813. [741] P.T. Reddy, P. Jaruga, G. Kirkali, G. Tuna, B.C. Nelson, M. Dizdaroglu. Identification and quantification of human DNA repair Protein NEIL1 by liquid chromatography/isotope-dilution tandem mass spectrometry, J. Proteome. Res. 12 (2013) 1049-1061. [742] G. Kirkali, P. Jaruga, P.T. Reddy, A. Tona, B.C. Nelson, M. Li, D.M. Wilson, III, M. Dizdaroglu. Identification and quantification of DNA repair protein apurinic/apyrimidinic endonuclease 1 (APE1) in human cells by liquid chromatography/isotope-dilution tandem mass spectrometry, PLoS. ONE. 8 (2013) e69894.

106

Figure Legends

Fig. 1. Structures of oxidatively induced DNA base lesions in DNA.

Fig. 2. Structures of the lesions derived from the sugar moiety of DNA.

Fig. 3. Tandem lesions identified in vivo [151,159,160,178-180].

Fig. 4. Levels of FapyAde and 8-OH-Gua in wt-, nth1‒/‒, ogg1‒/‒ and nth1‒/‒/ogg1‒/‒ mice

(data from [250]).

Fig. 5. Mispairings of oxidatively induced DNA lesions (redrawn from the data in

[297,349,356,358,367,370,399,430,434]).

Fig. 6. Left: Formation of C → T transition mutations from the mispairing of 5-OH-Cyt or 5-

OH-Ura with Ade [348,380,386,395-397]; Right: Formation of G → A transition mutations from the mispairing of S-cdG with Thy [297,430].

Fig. 7. Tumor incidence in neil1‒/‒, nth1‒/‒ and neil1‒/‒/nth1‒/‒ mice. Upper graphs: Lung

adenoma, adenocarcinoma. Lower graphs: Liver hepatocellular carcinoma, nodular

hyperplasia, severe dysplasia. Number (n) of animals: Males, neil1‒/‒, n = 25; nth1‒/‒, n = 52;

neil1‒/‒/nth1‒/‒, n = 43. Females, neil1‒/, n = 18; nth1‒/‒, n = 54; neil1‒/‒/nth1‒/‒, n = 29 (data

from [251]).

Fig. 8. Levels of R-cdA (A), S-cdA (B) and 8-OH-dG (C) in urine of control individuals (1)

and atherosclerosis patients (2) (data from [128]).

Fig. 9. Inhibition of the activity of human NEIL1 by small molecule inhibitors (data from

[738]).

107

guanine-derived products O O

N N NH NH O O N O O N N HN NHCHO N O N NH2 N NH2 HN H OH H2N HO H2N H O H O H2N N N H2N N NH2 H2N NH2 O H N NH2 N H H H H H H H H - , ------, , - - 8 hydroxyguanine 2 6 diamino 4 hydroxy 2,5 diamino 4H 2 2 4 triamino OH H OH H - - - - 5 formamidopyrimidine imidazol 4 one 5(2H) oxazolone (5'R)-8,5'-cyclo-2'-deoxyguanosine (5'S)-8,5'-cyclo-2'-deoxyguanosine

8-hydroxyguanine-derived products

H O N H O N O N NH N O H2N H N N H2N N H O H H H - spiroiminodihydantoin 5 guanidinohydantoin

adenine-derived products NH2 NH2

N N N N

NH2 NH2 NH2 N N N N N NHCHO N N N N HO H OH H H O HO O NH H H H H H N N H N 2 HO N N H H H H H H ------8 hydroxyadenine 4,6 diamino 5 form 2 hydroxyadenine OH H OH H ami o rimi ine d py d (5'R)-8,5'-cyclo-2'-deoxyadenosine (5'S)-8,5'-cyclo-2'-deoxyadenosine

cytosine-derived products

NH2 NH2 NH2 NH2 NH2 H H H OH N OH N OH N N OH N H OH H H OH H O N H O N O N O N H O N H H H H H H ------cytosine glycol 5 hydroxycytosine 5,6 dihydroxy 5 hydroxy 6 5,6 dihydrocytosine cytosine hydrocytosine

O O O O O O O H O H H OH H O O HN OH HN HN OH HN HN HN HN OH HN H OH H OH H O H H O N H N O N H O N OH O N O O N O N O O N H H H H H H H H H H ------, - uracil glycol 5 hydroxyuracil 5 hydroxy 6 5 hydroxy alloxan dialuric acid isodialuric acid 5 6 dihydrouracil hydrouracil hydantoin

thymine-derived products

O O O O O O CH3 CH3 CH OH CHO CH HN 2 HN HN 3 HN OH HN OH HN H H OH CH3 H O O N H O N H O N OH O N H N H H O N H H H H H H

- rox - - ro- m ne co - rox me - - orm urac - rox - - , - ro m ne 5 h y d y 6 hyd thy i gly l 5 ( h y d y thyl) 5 f yl il 5 hyd y 5 5 6 dihyd thy i thymine uracil methylhydantoin

Figure 1 sugar products released from DNA

CH HO CH HO CH 3 O 2 O 2 O O O O H H H O H H H O H H H H H H H

OH H H H OH H OH H

2,5-dideoxypentos-4-ulose 2,3-dideoxypentos-4-ulose 2-deoxypentos-4-ulose 2-deoxytetradialdose

sugar products bound to DNA as end group

CH3 ~ HO CH2 O O P O CH CH O H H 2 O 2 O H H H O O H O H H ~ O H H O ~H H P H O ~H P OH H O H P 2,5-dideoxypentos-4-ulose 2-deoxypentos-4-ulose 2-deoxypentos-4-ulose 2-deoxypentos-4-ulose

H ~

~ O C base O O O P O CH2 H H P O CH2 H H O H H O H H H O H H O ~H O O ~ H H H P P 2,3-dideoxypentos-4-ulose 2-deoxytetradialdose 3'-phosphoglycolate 5'-aldehyde

sugar products within DNA ~ ~ ~

P O CH2 P O CH2 P O CH2 O OH OH O H H H H COOH H H H H H O ~H O ~H O ~ O P P P

2-deoxypentos-4-ulose 2-deoxypentonic acid erythrose

Figure 2 O N O HN N HN O OH DNA H2N N N ~ DNA H2N N N ~ HN CH2 P O CH2 O O P O CH2 O NH2 H H H H2C O N NH H H H N H H CH2 O H H DNA H N O O H - - H N O HN CH CHO P O CH2 P O CH O 2 O H H H H thymine-tyrosine cross-link H H H H O ~ H O ~ H P DNA P DNA

Gua[8,5-Me]Thy Gua[8,5]Cyt

Figure 3 8-OH-Gua 4.0 FapyAde

3.0 DNA bases) 6 2.0

1.0 level (lesions/10

0.0

Figure 4 H N H H H N N O N H2N A N N N D H N N H N O D NA 8-OH-Gua•Ade

H H H N N H H H H O N O C N N N O H2N N H A N H N H N N N N D HN N H H N A C NH N NH D D O H N N N H 2 DNA A FapyGua•Ade FapyGua•Ade

H

H N N H H H O H O C N N H NH2 N H N N N N N H N N H N H D A N N A O N D NH N H N H N H N DNA DNA H

8-OH-Ade•Gua FapyAde•Ade

H H O N H N N H H N H N N N A N N D D NA O H

5-OH-Cyt•Ade (anionic imino tautomer)

D N A O H H CH3 D NA CH O H N H O 3 N N O H N O H H H H O N N N O N N H H N D O N H O H N H N A DNA O H A N D O N N H D H H A NH2 H H

S-cdG•Thy S-cdA•Thy

Figure 5 O

N NH

O NH2 N N NH2 OH OH H HN N C HO O H H O H H N O N H H H H OH H - - ' - , '- - '- 5 hydroxyuracil 5 hydroxycytosine (5 S) 8 5 cyclo 2 deoxyguanosine

G = C* *G=C

replication replication

A = C* *G =T

replication replication

A = T A=C* *G=T A = T

C → T G → A transition transition

Figure 6 males females 80 74.4% 50 41.4% 40 60

30 40 20

20 12% 10 cancer incidence (%) cancer incidence (%) cancer incidence 3.7% 1.9% 0% 0 0 -/- -/- -/- -/- -/- -/-

nth1 nth1 neil1 -/- /nth1 neil1 -/- /nth1

neil1 neil1

males females 50 46.5% 20 17%

40 15 13% 30 11.1% 10 20 16% 15.4% 5 10 12% 12% cancer incidence (%) cancer incidence (%) cancer incidence

0 0 -/- -/- -/- -/- -/- -/-

nth1 nth1 neil1 -/- /nth1 neil1 -/- /nth1

neil1 neil1

Figure 7 A 0.06 p < 0.0001 B 0.06 p < 0.0001 C 6 p = 0.0008

0.04 0.04 4 (nmol/mmol) (nmol/mmol) 0.02 (nmol/mmol) 0.02 2 R -cdA/creatinine S -cdA/creatinine 8-OH-dG/creatinine

0.00 0.00 0 1 2 1 2 1 2

Figure 8 NH2 O NHCHO NHCHO 75 N 160 HN NH H N 2 H2N N NH2 FapyAde FapyGua 60 120 DNA bases) DNA bases) 45 6 6 80 30

40 15 level (lesions/10 level (lesions/10

0 0

NEIL1 NEIL1 control control NEIL1NEIL1 + P2NEIL1 + P6NEIL1 + P7 + P8 NEIL1NEIL1 + P2NEIL1 + P6NEIL1 + P7 + P8 NEIL1NEIL1 + P11 + P19 NEIL1NEIL1 + P11 + P19 NEIL1 + DMSO NEIL1 + DMSO

S N

HN N HN S HN N N N N N N N N N

N N N N N N N N N H H H H H H

P2 (NCGC00188618) P6 (NCGC00188616) P7 (NCGC00188617) F

O N HN F NH HN O HN N N N N N N N N N F N N S N N N O O N N N H H H F H

P8 (NCGC00188619) P11 (NCGC00182914) P19 (MLS001126460)

Figure 9