Chemicals and Drugs Forming Reactive Quinone and Quinone Imine

Publication Date (Web): November 30, 2018. Copyright © 2018 American Chemical Society. Cite this:Chem. Res. Toxicol. XXXX, XXX, XXX-XXX ...
2 downloads 0 Views 2MB Size
Subscriber access provided by University of Winnipeg Library

Review

Chemicals and Drugs Forming Reactive Quinone and Quinone Imine Metabolites Ivana Klop#i#, and Marija Sollner Dolenc Chem. Res. Toxicol., Just Accepted Manuscript • DOI: 10.1021/acs.chemrestox.8b00213 • Publication Date (Web): 30 Nov 2018 Downloaded from http://pubs.acs.org on December 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1 2 3 4

Chemicals and Drugs Forming Reactive Quinone and Quinone Imine Metabolites

5

University of Ljubljana, Faculty of Pharmacy, Aškerčeva 7, 1000 Ljubljana, Slovenia

Ivana Klopčič and Marija Sollner Dolenc*

6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 1

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

45 46 47 48 49

50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 2

ACS Paragon Plus Environment

Page 2 of 84

Page 3 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121

Chemical Research in Toxicology

ABSTRACT: Quinones and quinone imines are highly reactive metabolites (RMs) able to induce dangerous effects in vivo. They are responsible for all kinds of toxicity: e.g. cytotoxicity, immunotoxicity and carcinogenesis. Furthermore, hepatotoxicity of chemicals/drugs in particular can be induced by quinone and quinone imine metabolites. According to their reactivity quinones and quinone imines react as Michael's acceptors with cell proteins and/or DNA and, in this way, cause damage to the cells. Quinones and quinone imines also have high redox potential and, due to their semiquinone radicals, are capable of redox cycling and forming reactive oxygen species (ROS). However, the presence of quinones and quinone imines structures in compounds is not always responsible for a toxic effect. The main question, therefore, is what are the main factors responsible for the toxicity of the chemicals and drugs that form RMs. For this reason, the presence of structural alerts and evidence for the formation of reactive quinones and quinone imines metabolites and their mechanisms of toxicity through cellular effects are discussed in this review, together with examples.

3

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158

Introduction Drug metabolism is associated with conversion of a xenobiotic to a less reactive hydrophilic metabolite that can be removed from the body. However, drug metabolism can lead, by another path, to chemically more reactive species, i.e. reactive metabolites (RMs).1 This process is, therefore, termed metabolic activation or bioactivation; it is known to have detrimental effects.1-3 Chemically, RMs are electrophiles (molecules that contain positive centers) or free radicals, and their toxicity can be mediated by covalent or noncovalent binding mechanisms.2 Electrophiles, which are inherently reactive, can be generated by the metabolism of chemically inert compounds. ‘Hard electrophiles’ (e.g., alkyl carbocation, carbonyl carbocation and nitrenium ion) and ‘soft’ electrophiles (e.g. Michael acceptors, quinone methide and quinone imine) are known. Factors such as the presence of a good leaving group, ring strain, polarization of a double bond by a Michael acceptor and the presence of electron withdrawing groups are crucial for the reactivity of electrophiles.4 Free radicals have an unpaired electron and form a covalent bond by reacting with another free radical, abstracting a hydrogen atom to form a new radical or abstracting an electron to generate a radical cation. Free radicals comprise those known as reactive oxygen species (ROS) and those that contain an additional nitrogen atom (e.g. nitric oxide (NO)). Despite their high reactivity their function is important in the vasculature, the nervous and immune systems, regulation of cell growth and in gene expression.5,6 Formation of RMs is also linked with an idiosyncratic adverse drug reaction (IDR), which is linked to life-threatening effects. The high reactivity of RMs may occasionally result in irreversible inhibition of the enzyme involved in their formation.7,8 Such toxicities cannot be disregarded by the pharmaceutical industry and a standard battery of toxicological assays has to be used routinely in a drug-discovery paradigm. RMs can react with nucleophilic centers of endogenous macromolecules to modify them covalently, causing toxicity. Metabolites can react with proteins leading to cell toxicity and immunogenicity.4,9 Reaction with nucleic acids can lead to changes in DNA and mutagenicity, teratogenicity or carcinogenicity.10-13

4

ACS Paragon Plus Environment

Page 4 of 84

Page 5 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

159 160 161 162 163 164 165 166 167 168 169

170 171 172 173 174 175 176

Chemical Research in Toxicology

Figure 1. Metabolic fate of xenobiotics in mammals.1 Quinones and quinone imines are highly redox active molecules and electrophiles, both these properties being crucial for their reactivity in biological systems. Furthermore, differences in chemical structure, in particular substituent effects, influence quinone toxicology.14 These compounds have a fully conjugated cyclic dione structure, such as that of benzoquinone (1, Figure 2), the prototypical member of the quinones.15

Figure 2. 1: p-Benzoquinone. Quinones as well as quinone imines are highly reactive organic chemicals and comprise a class of toxicological intermediates16 that interact alone or by generating ROS in biological systems to promote inflammatory reactions, reactivate immune cells, oxidize DNA, and in a such a way, inducing toxicity. They can be responsible for effects in vivo, including immunotoxicity, 5

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193

194 195 196 197 198 199 200 201 202 203 204 205 206

cytotoxicity and carcinogenesis.16 Two reactions of quinones and quinone imines are crucial for their activity: they reduce oxygen to reactive oxygen species, acting as prooxidants, and, as electrophiles, they form covalent bonds with tissue nucleophiles. Being Michael acceptors they can alkylate proteins and/or DNA, thereby being responsible for damaging cells. Thus, they can react with sulphur nucleophiles, such as GSH or cysteine residues on proteins, or with nucleophilic amino groups of proteins or DNA. Further, their activity can damage proteins, DNA, lipids and other cellular macromolecules through the formation of ROS that are responsible for oxidative stress by oxidizing these macromolecules. Further, it has been shown that the formation of 8-oxodeoxyguanosine is associated with ageing and carcinogenesis. In addition, protein kinase C and RAS can be activated by ROS.16 Regulatory proteins, such as protein tyrosine phosphatases, kelch-like ECH-associated protein 1, the regulatory protein for NF-E2-related factor 2, and the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase, are also possible targets for this type of reaction. Quinones and quinone imines also affect the pathways involved in cell signaling that protect against inflammatory responses and cell damage. These actions vary depending on the specific quinone and its concentration.17

Figure 3. Nucleophilic substitution of DNA and proteins and redox cycling of quinones. 2: Quinone; 3: Hydroquinone; 4: Semiquinone radical.16 Aromatic compounds such as phenols, hydroquinones, and catechols can be transformed to quinones by monooxygenase or peroxidase enzymes, metal ions and, in some cases, directly by molecular oxygen. 14, 18 Quinones are derivatives of their parent aromatic system: for example benzoquinones are derived from substituted benzene. For the effects of quinones on biological systems, two chemical properties are important that enable their reactivity, and their relative contribution to toxicity is influenced by substituent effects.14 They are considered as chemically reactive species because they are electrophilic compounds, but they can also undergo redox cycling and cause oxidative stress in cells (2, Figure 3).19 As a result of redox cycling, they can generate superoxide anion radicals.20

6

ACS Paragon Plus Environment

Page 6 of 84

Page 7 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241

Chemical Research in Toxicology

Oxygen intermediates (including the superoxide anion radical, hydrogen peroxide and hydroxyl radical) are termed reactive oxygen species (ROS). ROS can be generated by ultraviolet (UV) radiation or enzymatically. Free radicals have the potential of damaging every biological molecule present in the body. Molecules damaged in such a way can impair the functions of cells or even provoke their death, ultimately leading to different pathologies.21-23 For example, oxidation of proteins, caused by ROS, can generate protein hydroperoxides, which can generate additional radicals and may produce certain additional radicals, mainly on interaction with transition metal ions. Gradual accumulation of some oxidized proteins, that can occur with time, may contribute to the damages associated with age as well as to different pathological conditions.16 Lipids may be oxidized by ROS through generation of lipid peroxide-derived malondialdehyde DNA adducts.24 DNA is highly susceptible to damage by free radicals such as HO• since HO• very effectively attacks susceptible purines, generating 8-hydroxydeoxyguanosine (8-OHdG), 8-hydroxydeoxyadenosine, formamidopyrimidines and other less well characterized purine oxidative products. In addition, the reactivity of free radicals can also lead to the activation of poly-(ADP-ribose) synthetase, which may result in fragmentation of DNA and programmed cell death.21,23 From these data, it can be concluded that the compounds that form quinone and quinone imine metabolites may cause serious toxic effects. The main question is which factors affect the degree of toxicity of the chemicals and drugs that form quinones and quinone imines metabolites. The aim of this review is therefore represent the involvement of these metabolites in the toxicity of a variety of selected chemicals and drugs.

Selected chemicals and drugs- involvement of reactive quinone and quinone imine metabolites in their toxicity We focus on compounds with alerts in their chemical structure that have been associated with the formation quinone or quinonimine metabolites and with their role in compound toxicity (Table 1). In addition, not all RM-free compounds have guaranteed safety. Further, it is also well known that rare life-threatening reactions (such as hepatotoxicity, skin rashes and blood dyscracias) can be multifactorial in nature. The presence of structural alerts, evidence for quinone and quinone imine formation and their toxicity mechanisms through cellular effects will be discussed in this review. Table 1. Chemicals and drugs, their source, toxicity and the nature of the reactive intermediates. Chemical Source Toxicity Reactive Intermediate Solvent. Human Catechol, epoxide, Present in cigarette environmental hydroquinone, o smoke and car carcinogen. and pBenzene 19,25 30,32,33 exhaust. Produces bone benzoquinone. marrow damage in humans.26-28 Binds to proteins, 7

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Solvent.

Naphthalene



α-tocopherol analogue.

3,4‐dihydro‐6‐hydroxy‐2,2‐ dimethyl‐7‐methoxy‐1‐(2H)‐ benzopyran (CR‐6)

Ochratoxin A



Pentachlorophenol

DNA.29,30 Mechanism based inactivation of P450.31 GSH depletion and covalent binding to macromolecule s.34,35 Mechanism based inactivator of P450.36

Mycotoxin. Present Kidney toxicant in moldy cereals, in rodents.40 38,39 oats and beans. Possible carcinogen in humans (group B).41 Oxidative DNA, lipid and protein damage. Biocide. Liver toxicant Present in indoor in rodents.49,50 and outdoor air Endocrine pollution, food discruption and 47,48 chain. probable carcinogen in humans.51-57 Form macromolecule adducts, produce lipid peroxidation and DNA damage.42,58-62

8

ACS Paragon Plus Environment

Page 8 of 84

Quinone.35

Benzoquinone, hydroquinone, catechol, o-quinone and protein-thiol reactive metabolites.37 Tetrachlorobenzoquiquinone none.42-44,

OTQ, hydroquinone (OTHQ).45,46

Tetrachlorohydroqui none (Cl4HQ), tetrachlorocatechol (Cl4CAT)63, tetrachloro-1,4benzosemiquinone (Cl4-1,4-SQ) tetrachloro-1,2benzosemiquinone (Cl4-1,2-SQ), tetrachloro-1,4benzoquinone (Cl41,4-BQ) and tetrachloro-1,2benzoquinone (Cl41,2-BQ).42,58

Page 9 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

Pharmaceutical.

Liver toxicant Quinone imine.65 in humans.64 Binds to proteins and produces GSHdepletion.

Pharmaceutical.

Liver toxicant Benzoquinone in humans.66 imine.67,68

Pharmaceutical.

Liver toxicant O-quinone in humans.69 quinone imine.

Pharmaceutical.

Hepatotoxic, potentially genotoxic (group 3). Binds to proteins, GSH and DNA.70 Immunemediated toxicant in 71 humans. Binds to proteins and produces GSH depletion.72,73 Immunemediated toxicant, produces liver and kidney toxicity and hemolytic anemia.74



Lumiracoxib



Diclofenac



Tolcapone



Paracetamol

Pharmaceutical.



Amodiaquine

Pharmaceutical.

Nomifensine

or



9

ACS Paragon Plus Environment

Quinone imine.70

Quinone imine.72,73

N-hydroxylamine intermediate, quinone imine.75

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Pharmaceutical.



Lapatinib

Pharmaceutical.

Desatinib

Pharmaceutical.



Gefitinib

Pharmaceutical.



Erlotinib

Pharmaceutical.



Nefazodone

Pharmaceutical.

Liver toxicant in humans.76 Mechanism based on inactivation of P450.76 Induces inhibition of the bile salt export pump (BSEP).77 Liver toxicant in humans.78-80 Mechanism based inactivator of P450.78-80 Liver toxicant in humans.78-80 Mechanism based inactivator of P450.78-80 Exhibits BSEP inhibition.77 Liver toxicant in humans.78-80 Mechanism based inactivator of P450.78-80 Liver toxicant in humans.81,82 Mechanism based inactivator of P450.83 Mitochondrial toxicant and inhibitor of bile transport.84,85

Quinone imine.76

Produces hypersensitivity in humans.86

Epoxides,87,88 a quinone imine,89 and an ortho-quinone.90



Carbamazepine

10

Page 10 of 84

ACS Paragon Plus Environment

Quinone imine,78-80 imine-methide.78

Quinone imine.78-80

Quinone imine.78-80

Quinone imine and benzoquinone.81,82

Page 11 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

Pharmaceutical.

Responsible for Quinone imine.92 idiosyncratic drug toxicity 91 (IDT).

Pharmaceutical.

Not toxic to Quinone imine.82 humans at therapeutic dose.

Pharmaceutical.

Liver toxicity in Quinone imine94 and humans.93,94 epoxide.95

Pharmaceutical.

Not toxic humans therapeutic dose. Mechanism based inactivator P450.96 Not toxic humans therapeutic dose.98

in Catechol96 and at quinone.97

Not toxic humans therapeutic dose.

in Catechol and orthoat quinone.99

Atorvastatin



Aripiprazole



Trazodone



Paroxetine

Pharmaceutical.

Tadalafil

of in Catechol and orthoat quinone.98

Pharmaceutical.

Carvedilol

o-



11

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Pharmaceutical.

Mechanism based inactivator P450.100

Page 12 of 84

Raloxifene-6,7-oquinone.7,100,101 of



Raloxifene

Endogenous estrogen.



Estradiol

Forms DNA Quinone.102,103 adducts, produces carcinogenicity, cell transformation and mutagenicity.10 2,103

Quinone.105,106,108,109 Covalent a binding to DNA, in proteins.105,106. Oxidative stress in liver and tissues.104,107 or Neurotransmitter. Neurotoxicity.1 Quinone 10 110 semiquinone.



Bisphenol A

Dopamine

Environmental contaminant, monomer polycarbonate plastics.104



Pharmaceutical.

Doxorubicin

242 243 244 245 246 247 248 249 250 251 252

Cardiotoxocty Semiquinone or 111 and pulmonary hydroquinone. toxicity.111,112 Cytotoxicity, mutagenicity and DNA damage.



1. Chemicals and toxins that form quinone reactive metabolites 1.1. Benzene and role of its quinones metabolites in hematotoxicity and leukemogenicity Benzene is a ubitiquous human environmental carcinogen.113-116 It is widely used as an industrial chemical as a solvent and reagent for synthesis. It is a constituent of gasoline and cigarette smoke.19,25 The different ways of exposure to benzene affect a large proportion of the world population.117 Health concerns about benzene exposure arose when a recent study on shoe workers in China revealed that exposed people had lower numbers of granulocytes, lymphocytes and platelets in the peripheral blood than those in nonexposed people.19 Shorter exposure to benzene leads to early, reversible hematotoxicity, while longer exposure to high doses induces 12

ACS Paragon Plus Environment

Page 13 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273

274 275 276 277 278 279 280 281 282

Chemical Research in Toxicology

irreversible damage of bone marrow. The exposure to high-levels of benzene results in either the development of irreversible bone marrow aplasia (aplastic anemia) or, in case of patients who survive aplasia, in the appearance of dysplastic marrow (myelogenous leukemia).26,27,118 Although, the current U. S. occupational exposure limit is 1 ppm (parts per milion),19 there is still concern that prolonged/chronic exposure to low levels of benzene, due to environmental conditions, may be damaging.119 1.1.1. Benzene-reactive metabolites Exposure to benzene occurs primarily via inhalation. After absorption, it is biotransformed, through both phases I and II, in the liver.27,30,120 It has been shown that the toxicity of benzene can be reduced by partial hepatectomy and inhibition of benzene metabolism by toluene.121 The latter finding led to the suggestion that metabolism in the liver plays an important role in creating the toxic effect.122 Biotransformation involves the cytochrome P450, yielding benzene oxide in the first step, followed by converted into phenol, catechol, hydroquinone and 1,2,4trihydroxybenzene.19 This process requires the participation of the specific isoform of P450P450 2E1.123 Mice without P450 2E1 do not exhibit benzene-induced hematotoxicity.124 In humans, monohydroxylated benzene (phenol) is the major metabolite. There are other biotransformation pathways, such as ring opening, that lead to mucondialdehyde and muconic acid. It is assumed that the transformation of benzene to benzene oxide or its oxepin precedes the opening of the benzene ring.125

Figure 4. Metabolism of benzene. 5: Benzene; 6: Benzene oxide; 7: Dihydrodiol; 8: Phenol; 9: Catechol; 10: 1,2-Benzoquinone; 11: Hydroquinone; 12: 1,4-Benzoquinone; NQO1: NAD (P) H quinone oxidoreductase; MPO: myeloperoxidase; P450 2E1: cytochrome P450 2E1; EH: epoxide hydrolase.126 The hematotoxicity of benzene is mediated by its metabolites, such as phenol, catechol and hydroquinone. These are produced in liver121,127 and transported to the bone marrow128,129 where they are converted to biologically reactive intermediates such as p-benzoquinone by peroxidase-

13

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328

mediated oxidation. The bone marrow of humans, as well as of other animals, is therefore the primary target organ where toxic effects are expressed, due to, among other factors, its high myeloperoxidase activity and relatively low expression of NAD(P)H-quinone oxidoreductase (NQO). This combination results in rendering the two-electron transfer and inactivation of the pbenzoquinone back to the p-hydroquinone, and to a large number of rapidly dividing cells, in which cell damage is readily manifested. Further, bone marrow exhibits much lower levels of glutathione that is involved in xenobiotic protection.19,133 Reactive species are catechol, hydroquinone (HQ), and benzoquinone (BQ). Their toxicity is a result of either direct interaction with proteins (e.g. P450) and DNA, 30,32,33 or via oxygen activation.29,134 BQ thus affects tubulin by reacting with its thiol groups – that are strong nucleophiles. The consequence is the myelotoxicity of benzene that is caused by damage of the mitotic spindle. BQ can also reacts spontaneously with DNA, creating DNA adducts. The damaging effect of HQ on DNA is indirect, through the formation of ROS – specifically, hydroxyl radicals − that result from the dissociation of hydrogen peroxide and which is involved in 8-hydroxydeoxyguanosine formation.29,30 Oxidized forms of HQ damage P450 directly, rather than via oxygen activation of P450. Benzene metabolites destroy P450 in vitro, their ability decreasing in the order BQ > HQ > catechol > phenol. BQ or HQ damage P450 directly; this is not mediated by hydroxyl radical formation or by lipid peroxidation.31 In contrast, HQ and BQ protect P450 from destruction by radicals or the ROS that originate during the futile P450 cycle.32,135-137 HQ reacts spontaneously with oxygen, forming semiquinone radical and superoxide anion radical; O2•− may dismutate to H2O2. Furthermore, these two oxygen species form hydroxyl (HO•) radicals (the iron-catalyzed Haber-Weiss reaction generates HO• from H2O2 and O2•−) that are thought to be the most effective ROS in damaging DNA.134 Various benzene metabolites exhibit a variety of mechanisms of toxicity in hydroquinone-treated cells in which DNA adduct formation plays a role in the inhibition of cell differentiation but not in 1,2,4-benzenetriol-treated cells. 120 Myeloperoxidases convert hydroquinone to benzoquinone, requiring hydrogen peroxide in order to be catalytically active. Hydrogen peroxide is present at high concentrations in differentiated myeloid cells (polymorphonuclear leukocytes and monocytes) as well as in their progenitors.138140 Oxidation of hydroquinone by molecular oxygen or by copper ions results in the formation of benzoquinone and hydrogen peroxide, this reaction is a potential source of hydrogen peroxide.141-144 Myeloperoxidase is located in azurophilic granules; Cu/Zn SOD is cytoplasmic, and copper is found associated with DNA. Benzene metabolites, especially HQ and BQ can react with tubulin, topoisomerase II, histone and DNA.138 BQ can be reduced- by NAD(P)H: quinone oxidoreductase (NQO1)- back to hydroquinone which can be glucuronidated or sulphated for cell excretion. The alternative way to remove BQ is GSH which can react directly with BQ. Mutation of the gene for NQO1 leads to increased risk of benzene poisoning.145,146 GSH and NQO1 play roles in preventing hydroquinone-induced toxicity that vary according to the state of cell differentiation, cell type, and the species from which the bone marrow cells are derived.147 The in vivo method in mice of Lee et al.148 demonstrate the relative potency with which benzene and its metabolites inhibit erythropoiesis. Combining the hydroquinone plus muconaldehyde produces the greatest potency, while benzene has the lowest potential for toxicity. Also, all the benzene metabolites were highly effective when given in combination. Many of the benzene metabolites, except for phenol, were effective at lower iron uptake. The specific benzene metabolites that influence several cell-types and their different functions- are responsible for overall benzene toxicity. The effect of 1,2,4-benzenetriol needs more experiments on circulating blood cells and other hematopoietic tissue and further proofs of

14

ACS Paragon Plus Environment

Page 14 of 84

Page 15 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367

Chemical Research in Toxicology

its toxicity, since it was ineffective in preliminary studies. 1,2,4-benzenetriol can be isolated from mouse hepatocytes, suggesting a correlation with susceptibility to toxicity.120,149 The biological interactions of benzene are rather complex (5, Figure 4). Biotransformation of a structurally simple chemical is the main element in its hepato- and hemato-toxicity. Furthermore, benzene is one of the cases where strong evidence exists for RMs being responsible for its toxicity. The importance of understanding its toxicological actions in target cells underlies strategies for reducing and preventing them. 1.2. Naphthalene Naphthalene (NAP) has commonly been used in moth repellent products150, both as a solvent and a component of industrial products and waste materials.151 As a non-redox xenobiotic it can damage marine organisms and freshwater fish by producing ROS.151-153 NAP is genotoxic, and is an endocrine disruptor for fish154-156, a carcinogen in rats157 and, probably, in humans.158,159 Primary exposure of the general human population is through the environmental sources (indoor and outdoor air, water, soil, food) and in various work places (creosote impregnation, distillation of coal tar and naphthalene production, manufacture of refractories, graphite electrodes, aluminium and mothballs).157 Of the several human sites sensitive to the toxicity of naphthalene (e.g. eyes, lungs, liver, brain) the most prominent tissues associated with toxicity include the lens of the eye and the lungs.160 1.2.1. Bioactivation of naphthalene by P450 Metabolic activation of naphthalene starts by oxygene reduced nicotinic acid amide dinucleotide phosphate (NADPH)/P450-dependent mono-oxygenase enzymes (P450), which form very reactive, unstable stereoisomeric arene oxides, 1R,2S-epoxide and 1S,2R-epoxide (13, Figure 5).161,162 Epoxides can either be conjugated and excreted by kidneys or form 1-naphthol and 2naphthol (14 and 15, Figure 5), which can be eliminated directly as conjugates (16 and 17, Figure 5). Furthermore, the third metabolic pathway of epoxides is hydroxylation by epoxide hydrolase (EH) to form trans-1,2-dihydro-1,2-dihydroxynaphthalene (20, Figure 5). There are three metabolic pathways for dihydrodiol: in the first it becomes dehydrated, forming 2-naphthol (15, Figure 5); in the second it is oxidized to 1,2-dihydroxynaphthalene (21, Figure 5) by dihydrodiol dehydrogenases (aldo-keto-reductase (AKR)). Further, 1,2-naphthalenediol (21, Figure 5) can be metabolized to form 1,2-naphthoquinone (22, Figure 5) or the glucuronide/sulphate conjugate (23, Figure 5). The third metabolic pathway for the dihydrodiol (20, Figure 5) is glucuronic acid/sulphate conjugation (24, Figure 5) with potential subsequent dehydration (16, Figure 5). One more option for the metabolism of naphthalene is the formation of premercapturic acids and mercapturic acids (resulting from the glutathione conjugation), (25, Figure 5).163-168 1-Naphthol can be further oxidized to 1,4-dihydroxynaphthalene (18, Figure 5), which is oxidized to form 1,4-naphthoquinone (19, Figure 5).

15

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392

Figure 5. The main pathways and metabolites of naphthalene in mammals. 13: Naphthalene-1,2epoxide; 14: 1-naphthol; 15: 2-naphthol; 16: 1-naphthylglucuronide/sulphate; 17: 2naphthylglucuronide/sulphate; 18: 1,4-dihydroxynaphthalene; 19: 1,4-naphthoquinone; 20: trans1,2-dihydro-1,2-dihydroxynaphthalene; 21: 1,2-dihydroxynaphthalene; 22: 1,2-naphthoquinone; 23: 2-hydroxynaphthyl-1-glucuronide/sulphate; 24: trans-1,2-dihydro-2-hydroxynaphthyl-1glucuronide/sulphate; 25: naphthylmercapuric acids. UGT/ST: UDPglucuronyltransferase/sulphotransferase; GSH: glutathione-SH; R1: glucuronic acid/sulphate residue; R2: N-acetyl-L-cysteine residue, EH: epoxide hydrolase, AKR: aldo-keto-reductase.157 Reactive species (naphthoquinones) that are involved in naphthalene toxicity, could lead to increased production of ROS, such as superoxide anion and hydroxyl radical16,160,169-171, depleting glutathione and resulting in the formation of oxidative stress or in increased covalent adduct formation with proteins and DNA bases.172-176 1.3. CR-6 The 3,4-dihydro-6-hydroxy-2,2-dimethyl-7-methoxy-1-(2H)-benzopyran is a free radical scavenger and α-tocopherol analogue with potent inhibitory activity against lipid peroxidation in rat liver microsomes. It is neuro-protective in the disease where oxidative stress is a causative factor.37,177 CR-6 contains a nonsubstituted, activated aromatic position (C5) which explains its ability to act as a scavenger of nitric oxide and peroxynitrite by forming the corresponding 5nitroderivative and 1,4-benzoquinone.178 On oral administration, antioxidant CR-6 reaches the rat brain tissue and, after transient ischemia, protects the brain from re-perfusion injury.179 Furthermore, it may reduce infarct size caused by cerebral ischemia by preventing vessel hypertrophy and normalization of wall stress.180 As an NO scavenger and the protectant, CR-6 16

ACS Paragon Plus Environment

Page 16 of 84

Page 17 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422

Chemical Research in Toxicology

prevents glutamate neurotoxicity in cultures of cerebellar neurons.181 Antioxidants can be useful in a supportive therapy in diabetes: for example, CR-6, which is able to protect glutathione peroxidase (GPx) activity from glucose-induced hyperglycemia, effectively quenches nitrogen reactive species in Alloxan-induced experimental diabetes in mice.182 Reactive oxygen species (ROS) play a critical role in photoreceptor apoptosis while, on the other hand, CR-6 acts as a scavenger of ROS and reduces 661W photoreceptor apoptosis induced by nitric oxide donor sodium nitroprusside (SNP) by preventing activation of a pathway in which calpains play a key role.183 Sanvicens et al. highlight the relevance of CR-6, alone or in combination with other drugs, as a potential therapeutic strategy for the treatment of neurodegenerative diseases.184 Together, all the studies suggest that not only could CR-6 be a significant adjuvant therapy for retinal conditions, it could also be applied in other oxidative stress-induced disorders or neuropathies.177 1.3.1. Metabolic activation of CR-6 CR-6 is biotransformed in vitro via three different metabolic pathways in which the five metabolites (with the exception of o-quinone) are formed (Figure 6). In the first pathway, oxidative ring opening of the chromanol moiety initially produces benzoquinone (27a, Figure 6), which can undergo a non-enzymatically promoted cyclization to give the spiro compound (27b, Fgure 6). Alternatively, it can be reduced by P450 and NADP(H) to the corresponding hydroquinone (28, Figure 6). At the same time, the hydroquinone (28, Figure 6) could be reoxidized by P450 to benzoquinone (27a, Figure 6). The reactive oxygen species may also contribute to this reoxidation. Reactive oxygen species can be generated by redox cycling or by uncoupling of P450. In future work, it will be important to delineate whether the enzyme NAD (P) H quinone oxidoreductase (NQO1) plays a role in the reduction of CR-6 metabolite, 27abenzoquinone, since it was postulated that individuals lacking NQO1 may have a decreased capacity for protecting against cellular oxidative damage. The second oxidative pathway of CR-6 is via benzylic hydroxylation to compound 29 (Figure 6) by P450. The third is associated with the activity of P450, and generates catechol (30, Figure 6) through de-alkylation. The formation of the corresponding o-quinone derivative has also been identified, by HPLC (31, Figure 6).37

17

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452

Figure 6. P 450 mediated metabolic activation of CR-6. 26: CR-6 (3,4-dihydro-6-hydroxy-2,2dimethyl-7-methoxy-1-(2H)-benzopyran), 27a: 2-(3’-hydroxy-3’-methylbutyl-5-methoxy-1,4benzoquinone, 27b: oxaspiro[4.5]-2,2-dimethyl-8-methoxy-dec-8-ene-7,10-dione, 28: hydroquinone 2-(3’-hydroxy-3’-methylbutyl)-5-methoxyhydroquinone, 29: hydroxylated metabolite 3,4-dihydro-4,6-dihydroxy-2,2-dimethyl-7-methoxy-1(2H)-benzopyran, 30: catechol 3,4-dihydro-6,7-dihydroxy-2,2-dimethyl-1(2H)-benzopyran, 31: 3,4-dihydro-2,2-dimethyl1(2H)-benzopyran-6,7-dione.37 CR-6 antioxidant is an inhibitor of P450 isoenzyme activity. It showed inhibition of the 7ethoxyresorufin O-dealkylation activity of P450 1A and the 7-pentoxy-resorufin O-dealkylation activity of P450 2B in treated microsomes. The reactive metabolites, formed after CR-6 biotransformation, were stronger inhibitors of P450 2B than of P450 1A isoform. The stronger inhibition could be attributed either to the higher formation of potentially reactive metabolites for the P450 2B or the metabolites generated by P450 2B could bind more efficiently to this isoenzyme. Inactivation of P450-activities most probably occurs at the level of the P450 isoenzymes rather than at the level of the reductase. It is known that the activation of microsomal glutathione-S-transferases can be used to detect the thiol-reactive and potentially toxic chemicals that are formed,36 which happens as the result of oxidative and electrophilic compounds that react through a regulatory cysteine residue.185-188 CR-6 metabolites activate microsomal GSH transferases, since protein-thiol reactive metabolites are formed during P450-mediated biotransformation of CR-6. 37 The reactive metabolites formed are not mutagenic but are cytotoxic reactive intermediates. 1.4. Ochratoxin A Ochratoxin A (OTA) - (N-{[(3R)-5-chloro-8-hydroxy-3-methyl-1-oxo-7-isochromanyl] carbonyl}-3-phenyl-L-alanine) is a mycotoxine produced by some toxicogenic species of Aspergillus and Penicillium.189 Benzoquinone metabolites are also formed. It is a contaminant of moldy cereals, oats and has been extracted from some beans (coffee beans, peas, etc).38,39 OTA is a strong nephrotoxicant, neurotoxicant, immunotoxicant, mutagenic and teratogenic agent.190

18

ACS Paragon Plus Environment

Page 18 of 84

Page 19 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

453 454 455 456 457 458 459 460 461 462 463 464 465 466

Chemical Research in Toxicology

There is evidence that in rodents it causes kidney tumor.40 For human it is a possible carcinogen (group B), based on sufficient evidence for carcinogenicity in animals but inadequate evidence on humans.41 Because of OTA-mediated genotoxicity and its mechanism of carcinogenesis, the joint FAO/WHO Expert Committee has set a provisional tolerable weekly intake of OTA at 100 ng/kg body weight.191 There are two mechanisms of genotoxic activation: by forming a reactive oxygen species (ROS) that reacts with DNA to generate damage and promotes oxidative stress; the second by bioactivation and formation of covalent DNA adducts.192 Available data suggest that, for genotoxic effects, OTA must also have an indirect mechanism of action..193 1.4.1. Chemical structure of ochratoxins The chemical structures of ochratoxin A, B and C are shown below (32, Figure 7). Ochratoxin B is less toxic than ochratoxin A and does not inhibit protein biosynthesis in hepatoma cells.194 O

C

OR O

OH O O

N H

CH3 32

467 Ochratoxin A Ochratoxin B Ochratoxin C

468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490

X

X -Cl -H -Cl

R -H -H -CH2-CH3

Figure 7. 32: Ochratoxins. Ochratoxin A possesses a chlorine atom on C5 of the dihydro-methyl-isocoumarin ring system, and a phenolic OH group, which increases toxicity.196 This OH group can influence the secondary structure of ochratoxins by forming hydrogen bonds,197 and the C5-Cl atom of OTA plays an inportant role in OTB-dG formation.192 OTA is consists ofphenylalanine and dihydroisocumarin fragments which are coupled via an amide linkage. Splitting of this bond results in the formation of phenylalanine and ochratoxin α, which is not toxic.196,198 OTA is a weak acid (pKa = 7.1) that, under culture conditions, can form acid amides with tyrosine, glutamic acid, methionine, tryptophan, valine, serine, alanine and proline in place of phenylalanine. Some of them are very strong inhibitors of protein biosynthesis (for example alanine-ochratoxin A) while others are nonfunctional (for example proline-ochratoxin A).194 1.4.2. Oxidative stress and disruption of mitosis caused by OTA Oxidative stress plays a critical role in OTA carcinogenicity. The ROS formed cause oxidative damage to DNA.199-206 The latter has been detected indirectly by comet assay, by detecting the oxidized DNA bases207 and DNA fragmentation that are produced by the enzyme formamidopyrimidine-DNA-glycosylase (Fpg).200-206 Furthermore, OTA exposure is a source of both ROS and RNS,199 since it increases expression of nitric oxide synthase (iNOs) and stimulates protein nitration.208 ROS/RNA production increases levels of oxidative DNA, lipid and protein damage. OTA contains a phenolic ring system, and oxidative stress is mediated by phenols,209,210 which is responsible for thiol oxidation and antioxidant depletion and reduces GSH levels in mammalian

19

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514

515 516 517

cell lines.204 The second mechanism of indirect genotoxicity is disruption of mitosis for OTAmediated renal carcinogenesis. This is also the basis for making an additional validation for risk assessment.211 The group of researchers, categorize the toxin as a nonmutagenic, and propose that the mechanism of action (MOA) for indirect gentotoxicity is disruption of mitosis and chromosomal instability.211-214 OTA blocks the transition between metaphase/anaphase and leads to the formation of aberrant mitotic figures,212,213 inhibits microtubule assembly and blocks histone acetyltransferase (HAT) activity,214 which is a driving force in tumorigenesis. Furthermore, it is unlikely that disruption of mitosis plays a critical role in OTA- induced tumor formation, especially if karyomegaly is indicative of disruption.215,216 1.4.3. Direct genotoxicity of OTA Bioactivation of OTA generates reactive oxygen species (ROS) that attach covalently to DNA, generating DNA adducts that stimulate mutagenicity and renal carcinogenesis. In the presence of P450, OTA forms electrophilic tetrachlorobenzoquinone (TCBQ) that reacts covalently with sulfhydryl groups,42 2’-deoxyguanosine (dG),43 and other DNA bases44 to form benzetheno type adducts. Because it is well known that TCBQ forms covalent DNA adducts in rat liver, it is expected to play a key role in carcinogenesis.217 In the presence of peroxidase the electrophilic phenoxyl radical is formed by the oxidative pathway.218 The phenolic radical can react with the C8 site of dG to generate an oxygen- oxygen linked C8-OTA adduct,219 which is also formed when phenolic radical reacts with DNA (Figure 8).220 In the presence of the glutathione (GSH), the phenolic radical generates GS• that reacts with O2 to produce O2•−. This is converted with SOD to H2O2. Further H2O2 reacts in the Fenton reaction formed HO• which causes oxidative DNA damage. This pathway is important in the toxicity of all phenolic xenobiotics.209,210

Figure 8. Proposed pathways for the bioactivation of OTA. 33: OTA, 34: OTQ, 35: OTHQ, 36: OTA-GSH, 37: Phenoxyl radical, 38: Carbon-centered radical.192 20

ACS Paragon Plus Environment

Page 20 of 84

Page 21 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540

541 542 543 544 545 546 547 548 549 550

Chemical Research in Toxicology

Furthermore, OTA, in the presence of P450, generates the electrophilic quinone OTQ that reacts covalently with GSH to form the GSH conjugate (oxidative de-chlorination). The formed metabolite (OTQ) reacts with reducing agent (e.g. ascorbate) and it is reduced to hydroquinone N-{[(3R, S)-5,8-dihydroxy-3-methyl-1-oxo-3,4 dihydro-1H-isochromen-7-yl] carbonyl}-Lphenylalanine (OTHQ, 35).45,46 However, the reductive dehalogenation process leads to the formation of reactive aryl radicals, which intereact at the C8 site of purine bases to yield carbon (C)-linked DNA-adducts.221-223 Based on the OTA chemistry, we can suggest that the aryl radical can react with an H-donor to form OTB metabolite. Furthermore, this structure is appropriate for generating both reactive radical species (phenolic and aryl radicals) and an electrophilic quinone OTQ. 1.4.4. DNA adducts OTA forms guanine-DNA adducts shown in Figure 8. The first DNA-adduct was initially derived from the photoreactions of OTA in the presence of excess dG.224 They were also produced in the presence of Fe2+, Cu2+ and horseradish peroxidase.224 Furthermore, another theory was that the OTB-dG adduct was generated from attachment of the OTA with a phenoxyl radical.224 In fact, it was derived from direct aryl radical attachment to the dG, most likely with C8 site of dG,192 where Fe2+ acted as a reducing agent to cause reductive dehalogenation of OTA to afford the aryl radical species. Photolysis of OTA in the presence of dG gave a second adduct with the UV spectrum exposed suggesting loss of the phenolic H-atom. It is well know that photolysis of OTA generates the phenolic radical225 that reacts at the C8 site of dG (Figure 9).226, 227

Figure 9. Chemical structures of OTA-DNA adducts produced photochemically from the reaction of OTA or OTHQ with dG. 39: OTA-dG; 40: C-OTA-3΄-dGMP; 41: O-OTA-3΄dGMP.192 The final adduct was formed from the photoreaction of OTHQ in the presence of excess dG.228 There is a tendency of quinone electrophiles to react with dG to form benzetheno type adducts. The proposed structure for OTHQ-dG is confirmed by its mass observed by LC-MS.43,44,220 In conclusion, it is confirmed that OTA acts as a classical genotoxic carcinogen by bioactivation and covalent DNA addition.

21

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578

1.5. Pentachlorophenol (PCP) Pentachlorophenol (PCP) has been used as a pesticide, herbicide, algaecide, defoliant, wood preservative, germicide, fungicide, and moluscicide.48 PCP is a procarcinogen in rodents,50,229 and evidence suggests that it is an endocrine disruptor and probably carcinogen in humans,51-57 absorbed by the skin, lungs and gastrointestinal lining.230 Liver is the target organ of toxicity and carcinogenicity.49,229,231 Primary exposure is for the general human population, through the air and the food chain.47,48 Exposure to PCP has been reported in several studies including various age groups.232-236 High levels of PCP can lead to extremely high body temperature,237 which can cause a different kind of tissue injury and even death. PCP could have an immunotoxic function. High levels of PCP have, for a long time, been associated with malignant lymphoma and leukemia.238 Peripheral neuropathy and nerve damage are also described in people exposed to PCP.239 PCP induces oxidative stress and apoptosis, cell cycle arrest because of DNA damage (p53), mitogenic response (cyclin D1) and apoptosis (caspase 3).240 1.5.1. Metabolic activation of pentachlorophenol to quinones and conseqences One of the reasons why PCP induces liver tumor is its metabolism to chlorinated quinones in the liver59,241 by two proposed different pathways. The first is P450-mediated dechlorination, which gives tetrachlorohydroquinone (Cl4HQ) and tetrachlorocatechol (Cl4CAT),63 the latter being less prone to auto-oxidation than Cl4HQ.242 This auto-oxidation leads to subsequent oxidation of the hydroquinone to semiquinones, which could also proceed via an enzymatic pathway. Furthermore, these two metabolites are oxidized to semiquinones (i.e. tetrachloro-1,4benzosemiquinone (Cl4-1,4-SQ) and tetrachloro-1,2-benzosemiquinone (Cl4-1,2-SQ)), which are then oxidized to the respective quinones (i.e. tetrachloro-1,4-benzoquinone (Cl4-1,4-BQ) and tetrachloro-1,2-benzoquinone (Cl4-1,2-BQ)).42,58 The second pathway is direct cytochrome P450/peroxidase-mediated oxidation of PCP to (Cl4-1, 4-BQ), and includes loss of the chlorine anion (Figure 10).218,243,244

22

ACS Paragon Plus Environment

Page 22 of 84

Page 23 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600

Chemical Research in Toxicology

Figure 10. Proposed pathways for the PCP bioactivation to quinones and semiquinones. 42: PCP; 43: Cl4-HQ; 44: Cl4-CAT; 45: Cl4-1,4-SG; 46: Cl4-1,2-SG; 47: Cl4-1,4-BQ; 48: Cl4-1,2BQ.242 As quinones and semiquinones are electrophiles they form adducts with nucleophilic parts of macromolecules.42,58-61,245 Experiments on rats and mice show that Cl4-1, 4-BQ reacts with the sulfhydryl groups of cysteinyl residues in blood42 and with liver proteins245 to give mono-S- and multi-S-substituted adducts. Furthermore, Cl4-1,4,-SQ hydroxyl radicals, produced via metalindependent Fenton reactions,62 induce DNA damage. Furthermore, lipid peroxidation is an important factor in PCP-induced hepatic toxicity in mice because, under conditions of oxidative stress, lipid hydroperoxides can mediate the bioactivation of PCP to quinones or semiquinones, but much more than those associated with microsomal P450/NADPH.242 This is why lipid hydroperoxides may be predominantly important for the ultimate carcinogenicity of PCP.

2. Drugs that form quinone imine reactive metabolites 2.1. Lumiracoxib Lumiracoxib is an arylacetic acid derivative, used in the treatment of osteoarthritis, rheumatoid arthritis and acute pain. It is a carboxylic acid with weak acidic properties, which may be the reason for its pharmacokinetic and pharmacodynamic profile.246

23

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

601 602 603 604 605 606 607 608 609 610 611 612 613

614 615 616 617 618 619 620 621

Page 24 of 84

Figure 11. Structure of 49: 2-[(2-fluoro-6-chlorophenyl) amino]-5-methyl-benzene acetic acid. In humans, it is metabolized primarly via P450 enzymes.246 The major circulating metabolite of lumiracoxib in human is 4-hydroxylumiracoxib.64,65 It has recently been shown that 4hydroxylumiracoxib is bioactivated to a proposed quinone imine. N-acetylcysteine (NAC) trapped this quinine-imine metabolite by forming two NAC adducts.65 Furthermore, P450 and peroxidases bioactivated 4-hydroxylumiracoxib to a quinine-imine and several GSH-conjugated quinine-imine intermediates.64 In summary, bioactivation of lumiracoxib to quinine-imines may result in GSH-depletion, because it could interact up to four GSH molecules, bind covalently to proteins and trigger oxidative stress. Altogether, these risk factors lead to hepatotoxicity.64

Figure 12. Bioactivation of lumiracoxib in humans. Hydroxylumiracoxib; 52: Quinone imine; 53: GSH.64

50:

Lumiracoxib;

51:

4-

Structurally, lumiracoxib is related to diclofenac (54, Figure 13), the only differences being a methyl substituent on the 5-position and replacement of the dichloroaniline moiety of diclofenac by a fluorochloroaniline ring system (49, Figure 11).65

24

ACS Paragon Plus Environment

Page 25 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

622 623 624 625 626 627 628 629 630 631 632 633 634 635 636

Chemical Research in Toxicology

Figure 13. Structure of 54: Diclofenac. 2.2. Diclofenac Diclofenac is a nonsteroidal anti-inflammatory drug that has the same indications as selective COX-2 inhibitors. The reason for rare but potentially severe liver injury is metabolic activation of the drug.66 Diclofenac undergoes P450 2C9 and P450 3A4-catalyzed oxidation in humans, resulting in the formation of 4-hydroxydiclofenac and 5-hydroxydiclofenac, the products of which are also oxidized to reactive quinone imine intermediates. These react further with glutathione or microsomal proteins.67,68 Furthermore, glutathione adducts undergo sequential hydrolysis by γ-glutamyltranspeptidase and dipeptidases to give cysteinyl-glycine and cysteine derivatives.66 Ultimately, N-acetylation leads to the formation of the mercapturic acids, which are detected in the urine.66 We can conclude from this informationthat oxidative bioactivation of diclofenac in humans proceeds via bezoquinone imine intermediates (55, Figure 14).

25

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

637 638 639 640 641 642 643 644 645 646 647 648 649 650

Figure 14. Sheme for the P450-mediated bioactivation of diclofenac to quinone imines. 55: Diclofenac; 56: 4-Hydroxydiclofenac; 57: 5-Hydroxydiclofenac; 58: Diclofenac-2,5-quinone imine; 59: 5-OH-4-GS-diclofenac; 60: 5-OH-4-NAC-diclofenac; 61: Diclofenac-1,4- quinone imine; 62: 4-OH-3-GS-diclofenac; 63: 4-OH-3-NAC-diclofenac.66 2.3. Bioactivaton of drugs used for Parkinson′s disease Tolcapone is an orally active catechol-O-methyltransferase (COMT) inhibitor that is used in treating patients with Parkinson′s disease. It has been associated with several cases of liver injuries and, also, of fulminant hepatic failure.69,247 Catechol glucuronidation in position 3 is the major step for tolcapone metabolism in humans.247 The product of reduction of the 5-nitro group, namely the amine, can be excreted directly or after glucuronidation, sulphation or N-acetylation. Both the aniline and the N-acetylaniline derivatives are minor metabolites in human excreta.247 The next pathway is the oxidative hydroxylation of the methyl group to a primary alcohol and 26

ACS Paragon Plus Environment

Page 26 of 84

Page 27 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

651 652 653 654 655 656 657 658 659

660 661 662 663 664 665 666 667

Chemical Research in Toxicology

following oxidation to the carboxylic acid.247 The last step is 3-O methylation via COMT to the long-lived metabolite 3-OMT (73, Figure 15).247 The hepatotoxicity of tolcapone may be explained by the formation of reactive species (o-quinone or quinone imine species) through oxidation of the aniline and N-acetylaniline derivatives, which can form protein adducts with hepatic proteins.69 Bioactivation of tolcapone is thus the reason for its hepatotoxicity. Individuals who are carriers of mutations for the UDP-glucuronosyltransferase (UGT) 1A9 gene may exhibit more RM formation through the redox bioactivation pathway.248 Tolcapone is also a potent uncoupler of oxidative phosphorylation and an inhibitor of mitochondrial respiration.249

Figure 15. Metabolism of tolcapone in humans. 64: Tolcapone; 65: 3-O-β, D-glucuronide; 66: Amine derivative; 67: Quinone-imine; 68: Amine glucuronide; 69: Amine sulphate; 70: Quinone-imine; 71: Primary alcohol metabolite; 72: Carboxylic acid; 73: 3-Omethyltolcapone.247 2.4. Paracetamol

27

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

668 669 670 671 672 673 674 675 676 677 678 679 680

681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697

Paracetamol (PAR) is widely used as an over-the counter remedy to treat fever and pain. Some studies have reported PAR toxicity in reproduction and development, and proposed that they are linked to low sperm parameters and testicular germ cell cancer in humans.250,251 On the other hand, PAR has been shown to be an endocrine disruptor.252-254 It is a potentially genotoxic drug according to International Agency for Research on Cancer (IARC).255 Bioactivation of PAR by P450 enzymes (P450 1A1, 1A2 and 3A4) occurs at less than 5 % of the oral dose (74, Figure 16).256,257 Paracetamol is metabolized, to the fullest extent, to reactive N-acetyl-p-benzoquinone imine (NAPQI) by P450 2E1, leading to the formation of reactive N-acetyl-p-benzoquinone imine (NAPQI) metabolite, which can react with DNA, nucleophilic groups of proteins, and glutathione (GSH).70,258-260 Furthermore, increased levels of ROS lead to decreased levels of GSH or redox cycling of an electrophilic intermediate (NAPQI).261 QIs react, by a 1,4-Michael addition, with nucleophiles such as reduced GSH, protein thiols and nucleic bases.70

Figure 16. Paracetamol bioactivation to reactive species (quinone imine). 74: Paracetamol, 75: Paracetamol-o-glucironide, 76: Paracetamol sulphate, 77: N-acetyl-p-quinone imine (NAPQI), 78: Paracetamol 3-glutathione.262 2.5. Amodiaquine Amodiaquine (AQ) is a 4-aminoquinoline derivative used for treating malaria and rheumatoid diseases. Because of serious adverse side effects (agranulocytosis and liver damage), it is no longer used in frontline therapy for malaria. The detection of IgG antibodies to AQ in patients exposed to amodiaquine has shown that it is antigenic.71 The basis of the immune-mediated toxicity is a chemically reactive metabolite of AQ. Amodiaquine and its N-de-ethylated metabolite263 are metabolized by P450 and myeloperoxidase (MPO) to reactive quinine-imine species that can react covalently to cell proteins or GSH (80, Figure 17)72,73 Introduction of a fluorine group instead of hydroxyl into the antimalarial amodiaquine could prevent the process of oxidative bioactivation; this is evident for deOH-4-FAQ, which has antimalarial activity similar to that of amodiaquine but does not form a reactive metabolite in vivo.264

28

ACS Paragon Plus Environment

Page 28 of 84

Page 29 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726

Chemical Research in Toxicology

Figure 17. RM formation pathway for the basic drug (amodiaquine) and analogues (81, 82 and 83) incapable of forming RMs. 79: Amodiaquine; 80: Amodiaquine quinon imine.264 Furthermore, interchanging the 3- and 4-substituents in amodiaquine would lead to analogues 81, 82 and 83, (Figure 17) that are incapable of forming the quinone imine. These strategies have been used to modify 4-aminoquinoline amodiaquine by replacing the 4-hydroxy group with a hydrogen, fluorine or chlorine atom to give candidates with good safety profiles for further studies.265 2.6. Nomifensine Nomifensine, a tetrahydroisoquinoline derivative, is an antidepressant drug that was been withdrawn from market because of adverse side effects like hemolytic anemia and liver and kidney toxicity.74 In one study, antibodies and/or autoantibodies directed against the drug and/or metabolite- red blood cell conjugates were detected in patients who developed immune hemolytic anemia while receiving the drug. All the antibodies belonged to the IgG or IgM class or to both and were capable of activating complement.74 Bioactivation of aniline involves N-hydroxylation of the primary amine nitrogen, leading to formation of the N-hydroxylamine intermediate, which is converted nonenzymatically to the nitroso derivative which reacts with GSH to form an unstable mercaptal derivative that rearranges to a GSH-based sulfinamide conjugate.75 The sulfonyloxy group in the N-O sulfate conjugates of N-hydroxylamines is a very good leaving group that can be eliminated to generate a highly reactive nitrenium ion, which may lead to toxicity.75 The next step is N-glucuronidation, by which metabolites (N-glucuronides) can be hydrolyzed to the N-hydroxylamines which cause bladder and colon cancer. The second step involves the arene oxidation pathways, resulting in the formation quinone imine species that can be trapped by GSH (86, Figure 18).75 From this data it is again evident that the aniline and the arene groups are potential toxicophores, being capable of generating reactive quinone and quinone imine metabolites.75

29

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750

Figure 18. Bioactivation of the nomifensine; 84: Nomifensine; 85: Nomifensine nitrosoamine; 86: Quinone imine. 2.7. Tyrosine kinase inhibitors 2.7.1. Lapatinib Lapatinib is the first oral dual tyrosine kinase inhibitor of ErbB-1 and ErbB-2, approved by the US Food and Drug Administration in 2007 as a promising alternative and orally available drug for advanced metastatic breast cancer. However, a black-boxed warning for lapatinib was released after postmarketing surveillance and clinical trial reports of elevated liver enzymes and hepatotoxicity that developed several days to months after commencement of therapy. The exact mechanism by which lapatinib causes hepatotoxicity is unknown, but it has been shown to form reactive metabolites by P450-mediated oxidation.76 The drug is bioactivated by P450 3A4/5 to form O- and N-dealkylated metabolites that have the potential to be further oxidized in reactive quinine imine. Furthermore, these metabolites are trapped by GSH (89, Figure 19). Quinone imines are electrophilic and have a tendency to react with nucleophilic groups of cellular proteins, leading to toxicity.76 These electrophilic species may cause hepatotoxicity in two ways: the first is direct covalent modification of hepatic proteins, causing cell dysfunction, and the second is to form haptens which, in turn, trigger an immune response. It has been shown that the drug undergoes biotransformation to form quinine imine species that inhibit P450 3A4 irreversibly through mechanism-based inactivation, most probably through oxidation of its Odealkylated metabolite.76 Lapatinib is also shown to be hepatotoxic through the inhibition of the bile salt export pump (BSEP).77 The HLA-DQA1*02:01 allele carriers may also have an increased genetic risk associated with hepatotoxicity.266

30

ACS Paragon Plus Environment

Page 30 of 84

Page 31 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766

Chemical Research in Toxicology

Figure 19. The formation of reactive metabolites by P450 3A4. 87: Lapatinib; 88: O-dealkylated lapatinib; 89: Quinone imine; 90: LAPA-G1; 91: LAPA-G2.76 2.7.2. Dasatinib, gefitinib and erlotinib All these three drugs have been approved for the treatment of various type of cancer and all are P450 3A4 mechanism-based inactivators acting through a reactive intermediate, quinine imine.7880 Hepatotoxicity is as a possible side effect of treatment. Gefitinib and erlotinib are metabolized, through oxidative defluorination, to a hydroxyaniline metabolite that may undergo P450mediated two-electron oxidation to form a reactive quinine-imine intermediate (96 and 98, Figure 20).79,80 Dasatinib inactivation proceeds through hydroxylation at the para-position of the 2chloro-6-methylphenyl ring, by further oxidation giving the reactive quinine imine intermediates. A minor pathway is formation of a reactive imine methide.78 Furthermore, gefitinib exhibits BSEP inhibition with strong potency.77

31

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

767 768

769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786

Figure 20. Cytochrome P450 mediated bioactivation of the desatinib, gefitinib and erlotinib. 92: Dasatinib; 93: Imine methide; 94: Quinone imine; 95: Gefitinib; 96: Quinone imine; 97: Erlotinib; 98: Quinone imine. 2.8. Nefazodone After nefazodone was approved as an antidepressant, it has been linked, since 1994, with idiosyncratic adverse reactions including hepatobiliary dysfunction and cholestasis. We can find reports that nefazodonehepatotoxicity has even been associated with liver transplantation or death at therapeutic doses in the range of 200-400 mg/day. Signs of liver failure appear 1 to 8 months after starting treatment with this drug. Nefazodone has been withdrawn in many countries in the European Union and Canada. The exact mechanism of nefazodone toxicity is not known but there is evidence for formation of reactive metabolite.81,82 The drug undergoes a P450 3A4 catalyzed aromatic hydroxylation that occurs para to the piperazinyl nitrogen. It leads to phydroxynefazodone, a major circulating metabolite, para-hydroxyaniline analog and wellestablished toxicophore. Furthermore, metabolites are oxidatively bioactivated in human liver microsomes to electrophilic quinonoid intermediates that may play a role in nefazodone-induced hepatic necrosis (100 and 101, Figure 21)81,82 Nefazodone is a weak inhibitor of cytochrome

32

ACS Paragon Plus Environment

Page 32 of 84

Page 33 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

787 788 789 790 791 792 793

794 795 796 797 798 799 800 801

802 803

Chemical Research in Toxicology

P450 2D6, does not inhibit P450 1A2 but inhibits P450 3A4.83 Inactivation of mentioned isoenzymes occurs by covalent adduct formation with reactive species (quinone imine or pbenzoquinone intermediates) on the enzyme.81 Nefazodone′s nonlinear pharmacokinetics and clinical drug-drug interactions with P450 3A4 substrates also indicate in vivo mechanism-based inactivation of the isoenzyme.81,83 The drug also exhibits mitochondrial toxicity and is a potent inhibitor of bile transport, most likely through these reactive metabolites. .84,85

Figure 21. Bioactivation pathways for the antidepressant nefazodone. 99: Nefazodone; 100: Quinone imine; 101: Benzoquinone; 102: GSH.82 The hepatotoxicity of nefazodone is based on the particular chemical substructure, so that chemicals with the same mechanism of action, but different chemical structure, are not hepatotoxic (e.g. buspirone) (103, Figure 22).84

Figure 22. Structure of 103: Buspirone.

33

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827

2.9. Carbamazepine Since 1960, the aromatic anticonvulsant carbamazepine has been used for treating epileptic patients. Adverse hypersensitivity reactions have been described for carbamazepine, especially skin rash (including toxic epidermal necrolysis (Lyell΄s syndrome)). Among the undesirable effects that may be associated with hypersensitivity reactions are also fever, hepatitis, bone marrow toxicity, pneumonitis and pseudolymphoma. These symptoms has been validated to shown in 1 of 5000 patients.86 Although the exact mechanism of carbamazepine-mediated hypersensitivity is not known, the structures of metabolites show that multiple epoxides and cyclic peroxide are involved in the in vivo metabolism and formation of RM by man.267 Thirtytwo metabolites of carbamazepine have been detected in epileptic patient urine.268 An in vitro study has shown bioactivation of carbamazepine with the formation of cytotoxic, protein-reactive and stable metabolites in human liver microsomes (HLM). The major bioactivation pathways are N-glucuronidation with formation of the active carbamazepine-10,11-epoxide metabolite, and hydroxylation on the aromatic rings that lead to 2- and 3-hydroxycarbamazepine derivatives.269,270 Several of these metabolites of the drug have been proposed as reactive species. They could lead to cell death by covalent binding to cell macromolecules or by acting as haptens, causing idiosyncratic toxicity. Reactive species are formed through 10,11-epoxidation and 2,3arene oxidation, including epoxides (carbamazepine-2,3-epoxide and carbamazepine-10,11epoxide),87,88 a quinone imine metabolite derived from metabolism of major metabolite 2hydroxycarbamazepine,89 and an ortho-quinone metabolite that is formed from catechol 2,3dihydroxy carbamazepine (Figure 23).90 In vivo studies that include hapten formation, are compatible with the known fact that sera from carbamazepine-hypersensitive patients often contain antibodies that recognize P450 3A proteins.271,272

34

ACS Paragon Plus Environment

Page 34 of 84

Page 35 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848

Chemical Research in Toxicology

Figure 23. Bioactivation of the anticonvulsant drug carbamazepine to reactive species. 104: Carbamazepine; 105: Carbamazepine-10,11-epoxide; 106 and 107: 1:1 ratio; 108: Carbamazepine-2,3-epoxide; 109: 2-Hydroxycarbamazepine; 110: 3-Hydroxycarbamazepine; 111: 2-Hydroxyiminostilbene; 112: Quinone imine; 113: 2,3-dihydroxycarbamazepine; 114: oQuinone.90 2.10. Atorvastatin Atorvastatin is the 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase inhibitor and so well-established drug for treating hypercholesterolemia. Although such treatment reduces cardiovascular events and mortality rates, it also shows adverse effects, such as myopathy, that can progress to rhabdomyolysis and asymptomatic increase in hepatic transaminases.273 A major bio-activation pathway is P450 3A4-mediated monohydroxylation on the acetanilide structural alert with formation of the ortho- and para-hydroxyacetanilide metabolites, which are oxidized to reactive quinone imine species.92 These reactive metabolites have been shown to bind covalently to HLM, and are responsible for idiosyncratic drug toxicity.91 This type of toxicity is often very serious and appears as severe hepatotoxicity, agranulocytosis, neutropenia, Stevens-Johnsons syndrome, but has also been found late in drug development or in the postmarketing phase. 2.11. Aripiprazole 35

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

849 850 851 852 853 854 855 856 857 858 859 860 861 862

863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879

Since its introduction in 2003, aripiprazole has been used as an atypical antipsychotic. As with nefazodone, P450-mediated metabolism on the 2,3-dichlorophenyl ring in aripiprazole resulted in the formation of the corresponding circulating para-hydroxyaniline metabolite, which may have a causal role for the formation of reactive quinone imines, similar to that of nephazodone (116 and 117, Figure 24). However, this circulating metabolite does not form GSH conjugates, suggesting that the bioactivation pathway does not end in a manner analogous to that of nefazodone.82 Despite liability to bioactivation and chronic use of aripiprazole there have been no reports of idiosyncratic hepatotoxicity associated with this drug. The explanation lies in the dose levels and pharmacokinetics of the drug. The daily dose of aripiprazole is 10 to 30-fold lower (520 mg once per day) than that for nefazodone (200-400 mg once per day), which may reduce the total body burden to reactive metabolite exposure upon aripiprazole administration. There are several examples of structurally similar drugs with different safety profiles, and drugs dosed at greater levels are associated with idiosyncratic adverse drug reactions (ADRs).

Figure 24. Reactive metabolites of antipsychotic drug aripiprazole in HLM. 115: Aripiprazole; 116: Quinone imine; 117: Quinone imine. 2.12. Trazodone Trazodone has been in use as a second-generation non-tricyclic antidepressant. It is coprescribed with other antidepressants, because of its less anticholinergic and sedative side effects. In cases of CNS disorders it is not used as a frontline therapy, since it may cause rare hepatotoxicity274 that has been associated with RM formation (120 and 121, Figure 25).93,94 As in nefazodone, the 3-chlorophenylpiperazine ring in trazodone is P450 3A4-hydroxylated to give corresponding para-hydroxytrazodone metabolite. It has been detected in human urine as a major metabolite of the drug.95 The para-hydroxy-m-chlorophenylpiperazine portion, through two electron oxidation, produced a reactive quinone imine species, which was trapped with GSH in HLM incubations of trazodone.94 Furthermore, N-dealkylation of trazodone to form m-chlorophenylpiperazine was mediated by P450 3A4, which, like trazodone, was also bioactivated.94 The second step of trazodone bioactivation is the formation of epoxide, which has been confirmed by detection of a 36

ACS Paragon Plus Environment

Page 36 of 84

Page 37 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

880 881 882 883 884

885 886 887 888 889 890 891 892 893 894 895

Chemical Research in Toxicology

dihydrodiol metabolite of trazodone in human urine.95 Trazodone can thus form stable dihydrodiol or/and a GSH-conjugate derived from epoxide ring opening.81 It has been shown that trazodone collapses the mitochondrial membrane potential, and imposes oxidative stress leading to cell death to which could also contribute the above-described reactive metabolites.85

Figure 25. Bioactivation of the trazodone. 118: Trazodone; 119: 4-Hydroxytrazodone; 120: Quinone imine; 121: Quinone imine; 122: Epoxide; 123: Dihydrodiol.

3. Presence of methylene dioxyphenyl functional group in chemical: structural alert or not for forming quinones? 3.1. Paroxetine Paroxetine is a selective inhibitor of serotonin reuptake and antidepressant that contains the 1,3benzidioxole (methylenedioxyphenyl) structural alert and has been used for treating various CNS disorders. Drugs that are metabolized by cytochrome P450 2D6 enter into drug-drug interactions 37

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

896 897 898 899 900 901 902 903 904 905 906 907 908 909 910

911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926

(DDIs) when coadministrated with paroxetine, because of its potent inhibition of human liver microsomal P450 2D6 activity via, apparently, a competitive mechanism. Furthermore, paroxetine is metabolized through P450 2D6 mediated 1,3-benzidioxole ring cleavage to a catechol intermediate. This process is also a mechanism-based inactivation of the P450 isoenzyme.96 In vitro studies with [3H] paroxetine have shown covalent binding to human liver microsomes and to S-9 proteins, and characterization of glucuronide and sulfate metabolites of reactive quinone species (126, Figure 26).97 One of the most important detoxification pathways is nucleophilic substitution of reactive intermediates with GSH or S-adenosyl methionine. As in Figure 26, the paroxetine-catechol can be O-methylated by COMT or oxidized to the reactive oquinone metabolite, which is scavenged by GSH. Both pathways lead to a significant reduction in covalent binding to the cell macromolecules. Furthermore, in vitro covalent binding cannot necessarily be assumed to be predictive of toxicity. The latter is also dependent on the total the dose of the drug, the amount of chemically reactive intermediate and knowledge of the contribution of metabolism that would detoxify the reactive metabolite.97

Figure 26. Bioactivation/Detoxication pathways of paroxetine in human hepatic tissue. 124: Paroxetine; 125: Paroxetine-catechol; 126: o-Quinone; 127: Human metabolites of paroxetine. 3.2. Tadalafil The next example of drug which forms the benzoquinone metabolites and has also the 1,3benzidioxole alert is tadalafil, a potent reversible phosphodiesterase-5 (PDE5) inhibitor used for the treatment of erectile dysfunction. Tadalafil, like paroxetine, undergoes P450 3A4-catalyzed bioactivation, generating an electrophilic catechol/ortho quinone species and causing inactivation of the enzyme. However, therapeutic concentrations of tadalafil (10-20 mg) do not produce idiosyncratic toxicity and/or DDIs in erectile dysfunction treatment.98 3.3. Carvedilol Carvedilol, a new antihypertensive drug with combined α- and β-adrenergic receptor antagonism, is administered as a racemic mixture of the (R)-(+) and (S)-(-) enantiomers. Quinone imine metabolites are reactive species of carvedilol that are formed through N-glucuronidation and 38

ACS Paragon Plus Environment

Page 38 of 84

Page 39 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946

947 948 949 950 951 952 953 954

Chemical Research in Toxicology

P450 2D6-mediated oxidative biotransformation pathways. The formation of reactive metabolites is expected due to presence of an aniline and a dialkoxyaromatic structural alerts in the drug.99,275 Over 50 metabolites have been confirmed in the studies of carvedilol metabolism, including among themcatechol/ortho-quinone species and GSH conjugates.99 Carvedilol is thus biotransformed also by aromatic hydroxylation on the carbazole ring to hydroxyaniline type metabolites.275 In the treatment of hypertension it is used from 6.25 mg to the maximum recommended daily dose of 100 mg without side effects, despite the fact that it forms so many reactive metabolites. 3.4. Raloxifene Raloxifene is a selective ER modulator that is prescribed in the treatment of osteoporosis in postmenopausal women and for the chemoprevention of breast cancer. In vitro studies have shown that the drug undergoes P450 3A4-catalyzed bioactivation, forming reactive metabolites that can be trapped by GSH. These reactive species have the potential to alkylate cell macromolecules, triggering toxicity. The major alkylating species was the relatively stable raloxifene 6,7-o-quinone (130, Figure 27).7,100,101 At the same time, the reactive species are suicide inhibitors of P450 3A4.100 However, in vivo studies have shown that the main pathway to detoxification of the phenolic groups is glucuronidation in the gut. This kind of mechanism is primarily responsible for the lack of side effects of the drug administered in the low daily dose.276

Figure 27. Oxidation of raloxifene to reactive species in human hepatic tissue. 128: Raloxifene; 129: 7-Hydroxyraloxifene; 130: Raloxifene-6,7-o-quinone; 131: Raloxifene di-quinone methide. (modif.101) 3.5. Estradiol Metabolic activation of estrogens plays an important role in the etiology of breast and ovarian cancer through adducts and oxidized bases that form mutations of genes, and/or via expression of

39

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

955 956 957 958 959 960 961 962

963 964 965 966 967 968 969 970 971 972 973 974 975

specific genes that normally control cell growth and proliferation.277,278 The International Agency for Research on Cancer has classified estradiol as a carcinogenfor humans. Bioactivation of estradiol (E2) begins by P450 1A1 and P450 1B1 that give two hydroxylated metabolites 2hydroxyestradiol (2-OH-E2) and 4-hydroxyestradiol (4-OH-E2), respectively (134 and 137, Figure 28). Furthermore, the catechols 2-OH-E2 and 4-OH-E2, with help of peroxidases, form semiquinones, and these can be further oxidized to quinones. The catechol-O-methyltransferase (COMT) defends organisms against genotoxic metabolites by methylation of catechol estrogens.279

Figure 28. Major metabolic pathways of the estradiol. 132: Estradiol, 133: Estrone, 134: 2Hydroxyestradiol, 135: Estradiol-2,3-semiquinone, 136: Estradiol-2,3-quinone, 137: 4Hydroxyestradiol, 138: Estradiol-3,4-semiquinone, 139: Estradiol-3,4-quinone. 17αHydroxysteroid dehydrogenase, 17α-HSD.279 3.6. Bisphenol A Bisphenol A (BPA) is industrial raw material and thus also an environmental contaminant, widely used as a monomer in polycarbonate plastic and epoxy resins.104,280,281 Humans are exposed to BPA through various products such as packaging for food and water, cans, thermal paper and dental sealants.282-284 BPA is an endocrine disruptor and may causes DNA damage as the result of its reactive metabolites that can form covalent adducts with nucleophilic macromolecules and/or produce oxidative stress (140, Figure 29).106 Exposure is also associated 40

ACS Paragon Plus Environment

Page 40 of 84

Page 41 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

976 977 978 979 980 981 982 983

984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002

Chemical Research in Toxicology

with inflammation and/or oxidative damage in brain, testes, kidney and sperm, and peroxidation reactions generated by BPA.285-288 Furthermore, it can produce harmful effects in liver of rats and mice, and adverse effects in pregnancy.104,289 The main way for decreasing toxicity is conjugation reactions that result in two main products, glucuronide and sulphate.290-296 The oxidized product of BPA is catechol analogue, which may act via semiquinone and/or quinone products to produce ROS and oxidative stress (141, Figure 29). BPA o-quinone metabolite could be also generated by polyphenol oxidase and tyrosinase, indicating that BPA is a substrate for many enzymes that allow the oxidation of both endogenous and exogenous compounds in organisms.297,298

Figure 29. Metabolic pathways of the bisphenol A. 140: Bisphenol A, 141: Bisphenol A catechol, 142: Bisphenol A quinone, 143: Bisphenol A ol, 144: Bisphenol A carboxylic acid, 145: 4-Methyl-2,4-bis-(p-hydroxyphenyl)pent-1-ene, 146: Bisphenol A glucuronide, 147: Bisphenol A sulfate.280 3.7. Dopamine Dopamine (DA), a central nervous system neurotransmitter, present in high concentrations in the hypothalamus is also an intermediate in the synthesis of both norepinephrine and epinephrine. It exhibits inherent toxic potential. Metabolic activation of DA may play a role in pathological processes related with schizophrenia, addiction, and Parkinson’s disease.299 DA can form reactive species via two pathways. One is production of ROS, partially by the enzyme monoamine oxidase (151, Figure 30).300 The second is spontaneous or, in the presence of transition metal ions (iron or manganese), oxidation to the reactive DA quinone.301,302 Additionally, a quinone moiety can be formed enzymatically by tyrosinase, prostaglandin H synthase and xanthine oxidase (150, Figure 30).303-307 Once the quinone is formed, it has cytotoxic and genotoxic potential in generating ROS, which play an important role in degenerative brain diseases.

41

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029

Figure 30. Metabolic pathways of dopamine. 148: Dopamine, 149: Dopamine o-semiquinone, 150: Dopamine quinone, 151: 3,4-dihydroxyphenylacetic acid (DOPAC) which also forms quinones upon oxidation, MAO: Monoamine oxidase, DBH: Dopamine beta-hydroxylase, NE: Norepinephrine, PNMT: Phenylethanolamine N-methyltransferase, EPI: Epinephrine.299 3.8. Doxorubicin Doxorubicin, a clinically active anticancer drug for treating solid and hematological malignancies, is used for inhibition and/or interference with topoisomerase II. However, the use of doxorubicin is limited by its accumulation in the heart, primarily in mitochondria, and dosedependent irreversible cardiomyopathy. Since it has a quinone fragment in its structure, it can interact with complex I in the respiratory chain and generate ROS. As a result, reduced ATP production and the activation of oxidative stress in the cell occur. Furthermore, doxorubicin can cause DNA cleavage and form double strand breaks that lead to cell death.308-310 Besides the noted mechanisms of action, bioactivation of doxorubicin may be important for its mode of action, as it produces reactive species that bind covalently, or induce damage, to various macromolecules.311 Doxorubicin contains a quinone-hydroquinone moiety, which is mainly reduced by P450 reductase,312 via one-electron reduction, to form a semiquinone radical (153, Figure 31). This unstable semiquinone radical, in the presence of molecular oxygen, undergoes an oxidation-reduction cycle to form superoxide radical and the hydroxyl radical (via further Fenton reaction, Figure 31), a highly reactive oxidant.312-316 However, one-electron reduction is followed by generation of highly reactive covalent binding species 154 and 155 that can alkylate cell macromolecules.317-320 A few studies have also confirmed DNA adduct formation.321-325 We can conclude that both the bioactivation and the quinone structure of doxorubicin cause the formation of ROS and the interaction of doxorubicin intermediates with macromolecules in the cell- mechanisms of its pharmacological as well as toxicological action.

42

ACS Paragon Plus Environment

Page 42 of 84

Page 43 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054

Chemical Research in Toxicology

Figure 31. Bioactivation of the doxorubicin. 152: Doxorubicin, 153: Semiquinone, 154: C7-Free radical, 155: C7-Quinone methide, 156: DNA adduct, R: CH2-OH.311

Conclusion and future directions All the presented chemicals and drugs exhibit their toxicity as a result of biooxidation. Most of them produce liver toxicity in humans; some were carcinogens while some presented drugs were not toxic in therapeutical doses, although they produce reactive species. The target organ toxicity is often driven by the tissue-specific expression of the P450 enzymes and peroxidases responsible for the bioactivation of the drug or chemical (example: bone marrow is the target organ of benzene toxicity because of the high myeloperoxidase activity responsible for the oxidation). Quinones and quinone imines are considered chemically reactive species, because they are electrophilic compounds, but they can also undergo redox cycling and cause oxidoreductive stress in cells. There are many other factors that impact on the overall toxicity of the discussed chemicals and drugs. First, the dose level and pharmacokinetics of the drug or chemical can influence its toxicity. For example lower daily dose levels of the drug may reduce the total body burden to reactive metabolite exposure upon drug administration. This is important in the case of aripiprazole which daily dose is 10 to 30-fold lower than the daily dose of nefazodone. Secondly, the compounds generally have several metabolic pathways and, depending on what prevails, produce more or less highly reactive metabolites, including quinones and quinone imines. The major pathway of detoxication of the reactive metabolites in humans is glucuronidation or reaction with glutathione, that usually leads in a less toxic metabolites. This explains why raloxifene is a safe drug despite the fact that it also undergoes P450 3A4-mediated

43

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070

1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085

oxidation forming stable alkylating reactive species (raloxifene 6,7-o-quinone). Thirdly, the overall toxicity of the compound on the specific cell targets is influenced by the reactivity of all its metabolites. Some of the reactive metabolites, such as benzene, gefitinib, dasatinib, lapatinib, nefazodone, paroxetine and raloxifene metabolites, react with enzyme active site causing immediate enzyme inactivation. The other reactive metabolites migrate from the site of formation and bind to proteins, DNA or produce oxidative DNA, lipid and protein damage. In order to understand these processes, the theory of “hard” and “soft“ electrophiles should be taken into account. GSH also plays an important role in modulating the toxicity of quinones and quinone imines as noted previously. Reactive intermediates react with GSH in the presence of glutathione-Stransferases to form GSH conjugates, which are then excreted by kidneys. These reactions protect cells against the harmful effects of reactive intermediates. GSH depletion consistently enhanced the toxicity of these compounds. The precise mechanism of enhancement has not been investigated. Whether GSH conjugates plays a role in the toxicity of chemicals and drugs, form quinone and quinone imine species, is a question that needs further investigation.

Figure 32. Mechanism of action for quinones and quinone imines. It has been shown that quinones and quinone imines play roles in mediating the toxic effects of amino and/or hydroxylated aromatic compounds. These electrophilic and redox active species can cause damage within cells by various pathways (Figure 32). It is well known that the molecular structure influences the location and type of cellular nucleophiles targeted. Oxidative enzymes, metal ions, and in some cases molecular oxygen can catalyze the formation of quinoids that interact with cell nucleophiles (GSH, proteins and DNA) to promote inflammatory, antiinflammatory, and anticancer actions and so induce toxicity. In addition, the formation of ROS, contributes to the cytotoxic properties of the parent compounds. Redox cycling generates reactive oxygen species that cause lipid peroxidation, depletion of reducing equivalents, oxidation of DNA and DNA strand breaks. Quinones and quinone imines could affect various cell signaling pathways that protect against inflammatory responses and cell damage. Further, ROS can activate signaling pathways, which include protein kinase C and RAS. DNA binding occurs, such as 44

ACS Paragon Plus Environment

Page 44 of 84

Page 45 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116

Chemical Research in Toxicology

formation of 8-oxodeoxyguanosine, which has been associated with ageing and carcinogenesis. For example: OTA acts as a classical genotoxic carcinogen by bioactivation and covalent DNA adduction. In addition, alkylation and relationships to cytotoxicity are dependent on the chemical structure and the cellular environment in which they are formed. Future work should investigate carcinogenic and cytotoxic metabolites such as quinones and quinone imines of aromatic amines and hydroxylated aromatics to show their molecular targets as well as mechanism(s) of action, but in the meantime in vitro safety prediction will remain a critical topic for further investigation and debate. AUTHOR INFORMATION Corresponding Author * Tel.: +386-1-476-9572. Fax: +386-1-425-8031. E-mail address: [email protected]. Funding One of us (I.K.) would like to acknowledge AdFutura foundation for awarded scholarship. This work was financed by program group “Medicinal Chemistry: Drug Design, Synthesis and Evaluation of the Drugs” (program code P1-0208). Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS The authors thank Dr. Roger Pain and Dr. Janez Mavri for their critical reading of the manuscript and linguistic corrections.

REFERENCES

1117 1118

(1) Attia, S. M. (2010) Deleterious Effects of Reactive Metabolites. Oxid. Med. Cell. Longev. 3,

1119

238–253.

1120

(2) Park, B. K., Kitteringham, N. R., Maggs, J. L., Pirmohamed, M., and Williams, D. P. (2005)

1121

The role of metabolic activation in drug-induced hepatotoxicity. Annu. Rev. Pharmacol. Toxicol.

1122

45, 177–202.

45

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1123

(3) Park, B. K., Laverty, H., Srivastava, A., Antoine, D. J., Naisbitt, D., and Williams, D. P.

1124

(2011) Drug bioactivation and protein adduct formation in the pathogenesis of drug-induced

1125

toxicity. Chem. Biol. Interact. 192, 30–36.

1126

(4) Uetrecht, J. (2007) Idiosyncratic drug reactions: current understanding. Annu. Rev.

1127

Pharmacol. Toxicol. 47, 513–539.

1128

(5) Vanhoutte, P. M. (2009) How we learned to say no. Arterioscler. Thromb. Vasc. Biol. 29,

1129

1156–1160.

1130

(6) Bedard, K., and Krause, K.-H. (2007) The NOX family of ROS-generating NADPH oxidases:

1131

physiology and pathophysiology. Physiol. Rev. 87, 245–313.

1132

(7) Chen, Q., Ngui, J. S., Doss, G. A., Wang, R. W., Cai, X., DiNinno, F. P., Blizzard, T. A.,

1133

Hammond, M. L., Stearns, R. A., Evans, D. C., Baillie, T. A., and Tang, W. (2002) Cytochrome

1134

P450 3A4-mediated bioactivation of raloxifene: irreversible enzyme inhibition and thiol adduct

1135

formation. Chem. Res. Toxicol. 15, 907–914.

1136

(8) Kalgutkar, A. S., Dalvie, D. K., Aubrecht, J., Smith, E. B., Coffing, S. L., Cheung, J. R.,

1137

Vage, C., Lame, M. E., Chiang, P., McClure, K. F., Maurer, T. S., Coelho, R. V., Soliman, V. F.,

1138

and Schildknegt, K. (2007) Genotoxicity of 2-(3-chlorobenzyloxy)-6-(piperazinyl)pyrazine, a

1139

novel 5-hydroxytryptamine2c receptor agonist for the treatment of obesity: role of metabolic

1140

activation. Drug Metab. Dispos. 35, 848–858.

1141

(9) Ikehata, K., Duzhak, T. G., Galeva, N. A., Ji, T., Koen, Y. M., and Hanzlik, R. P. (2008)

1142

Protein targets of reactive metabolites of thiobenzamide in rat liver in vivo. Chem. Res. Toxicol.

1143

21, 1432–1442.

1144

(10) Amacher, D. E. (2006) Reactive intermediates and the pathogenesis of adverse drug

1145

reactions: the toxicology perspective. Curr. Drug Metab. 7, 219–229.

46

ACS Paragon Plus Environment

Page 46 of 84

Page 47 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1146

(11) Wells, P. G., McCallum, G. P., Chen, C. S., Henderson, J. T., Lee, C. J. J., Perstin, J.,

1147

Preston, T. J., Wiley, M. J., and Wong, A. W. (2009) Oxidative stress in developmental origins of

1148

disease: teratogenesis, neurodevelopmental deficits, and cancer. Toxicol. Sci. 108, 4–18.

1149

(12) Wells, P. G., Lee, C. J. J., McCallum, G. P., Perstin, J., and Harper, P. A. (2010) Receptor

1150

and reactive intermediate-mediated mechanisms of teratogenesis. Handb. Exp. Pharmacol. 196,

1151

131–162.

1152

(13) Skipper, P. L., Kim, M. Y., Sun, H.-L. P., Wogan, G. N., and Tannenbaum, S. R. (2010)

1153

Monocyclic aromatic amines as potential human carcinogens: old is new again. Carcinogenesis

1154

31, 50–58.

1155

(14) Monks, T. J., and Jones, D. C. (2002) The metabolism and toxicity of quinones,

1156

quinonimines, quinone methides, and quinone-thioethers. Curr. Drug Metab. 3, 425–438.

1157

(15) McNaught, A. D., and Wilkinson, A. (1997) Compendium of Chemical Terminology. Wiley-

1158

Blackwell.

1159

(16) Bolton, J. L., Trush, M. A., Penning, T. M., Dryhurst, G., and Monks, T. J. (2000) Role of

1160

quinones in toxicology†. Chem. Res. Toxicol. 13, 135–160.

1161

(17) Kumagai, Y., Shinkai, Y., Miura, T., and Cho, A. K. (2012) The chemical biology of

1162

naphthoquinones and its environmental implications. Annu. Rev. Pharmacol. Toxicol. 52, 221–

1163

247.

1164

(18) Monks, T. J., Hanzlik, R. P., Cohen, G. M., Ross, D., and Graham, D. G. (1992) Quinone

1165

chemistry and toxicity. Toxicol. Appl. Pharm. 112, 2–16.

1166

(19) Boelsterli, U. A. (2003) Mechanic Toxicology. The molecular basis of how chemicals

1167

disrupt biological target. CRC Press.

1168

(20) O'Brien, P. J. (1991) Molecular mechanisms of quinone cytotoxicity. Chem. Biol. Interact.

1169

80, 1–41. 47

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 48 of 84

1170

(21) Devasagayam, T. P. A., Tilak, J. C., Boloor, K. K., Sane, K. S., Ghaskadbi, S. S., and Lele,

1171

R. D. (2004) Free radicals and antioxidants in human health: current status and future prospects.

1172

J. Assoc. Physicians. India 52, 794–804.

1173

(22) Valko, M., Leibfritz, D., Moncol, J., Cronin, M. T. D., Mazur, M., and Telser, J. (2007) Free

1174

radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem.

1175

Cell Biol. 39, 44–84.

1176

(23) Fang, Y.-Z., Yang, S., and Wu, G. (2002) Free radicals, antioxidants, and nutrition. Nutrition

1177

18, 872–879.

1178

(24) Marnett, L. J. (1999) Lipid peroxidation-DNA damage by malondialdehyde. Mutat. Res.

1179

424, 83–95.

1180

(25) Wallace, L. (1996) Environmental exposure to benzene: an update. Environ. Health

1181

Perspect. 104, 1129–1136.

1182

(26) Aksoy, M., Dinçol, K., Akgün, T., Erdem, S., and Dinçol, G. (1971) Haematological effects

1183

of chronic benzene poisoning in 217 workers. Br. J. Ind. Med. 28, 296–302.

1184

(27) Snyder, R., and Kocsis, J. J. (1975) Current concepts of chronic benzene toxicity. CRC Crit.

1185

Rev. Toxicol. 3, 265–288.

1186

(28) Infante, P. F., Rinsky, R. A., Wagoner, J. K., and Young, R. J. (1977) Leukaemia in benzene

1187

workers. Lancet 2, 76–78.

1188

(29) Leanderson, P., and Tagesson, C. (1990) Cigarette smoke-induced DNA-damage: role of

1189

hydroquinone

1190

hydroxydeoxyguanosine. Chem. Biol. Interact. 75, 71–81.

1191

(30) Snyder, R., Witz, G., and Goldstein, B. D. (1993) The toxicology of benzene. Environ.

1192

Health Perspect. 100, 293–306.

and

catechol

in

the

formation

48

of

the

ACS Paragon Plus Environment

oxidative

DNA-adduct,

8-

Page 49 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1193

(31) Gut, I., Nedelcheva, V., Soucek, P., Stopka, P., and Tichavská, B. (1996) Cytochromes P450

1194

in benzene metabolism and involvement of their metabolites and reactive oxygen species in

1195

toxicity. Environ. Health Perspect. 104, 1211–1218.

1196

(32) Soucek, P., Filipcova, B., and Gut, I. (1994) Cytochrome P450 destruction and radical

1197

scavenging by benzene and its metabolites. Evidence for the key role of quinones. Biochem.

1198

Pharmacol. 47, 2233–2242.

1199

(33) Irons, R. D. (1985) Quinones as toxic metabolites of benzene. J. Toxicol. Environ. Health

1200

16, 673–678.

1201

(34) Warren, D. L., Brown, D. L., and Buckpitt, A. R. (1982) Evidence for cytochrome P-450

1202

mediated metabolism in the bronchiolar damage by naphthalene. Chem. Biol. Interact. 40, 287–

1203

303.

1204

(35) Zheng, J., Cho, M., Jones, A. D., and Hammock, B. D. (1997) Evidence of quinone

1205

metabolites of naphthalene covalently bound to sulfur nucleophiles of proteins of murine Clara

1206

cells after exposure to naphthalene. Chem. Res. Toxicol. 10, 1008–1014.

1207

(36) Onderwater, R. C., Commandeur, J. N., Menge, W. M., and Vermeulen, N. P. (1999)

1208

Activation of microsomal glutathione S-transferase and inhibition of cytochrome P450 1A1

1209

activity as a model system for detecting protein alkylation by thiourea-containing compounds in

1210

rat liver microsomes. Chem. Res. Toxicol. 12, 396–402.

1211

(37) Yenes, S., Commandeur, J. N. M., Vermeulen, N. P. E., and Messeguer, A. (2004) In vitro

1212

biotransformation of 3,4-dihydro-6-hydroxy-2,2-dimethyl-7-methoxy-1(2H)-benzopyran (CR-6),

1213

a potent lipid peroxidation inhibitor and nitric oxide scavenger, in rat liver microsomes. Chem.

1214

Res. Toxicol. 17, 904–913.

1215

(38) Pohland, A. E., Nesheim, S., and Friedman, L. Ochratoxin A: a review (technical report).

1216

Pure Appl. Chem. 64, 1029-1046. 49

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 50 of 84

1217

(39) Jørgensen, K. (2005) Occurrence of ochratoxin A in commodities and processed food-a

1218

review of EU occurrence data. Food Addit. Contam. 22, 26–30.

1219

(40) (1989) Toxicology and carcinogenesis studies of ochratoxin A (CAS No. 303-47-9) in

1220

F344/N rats (gavage studies). Natl. Toxicol. Program 358, 1-142.

1221

(41) International Agency for Research on Cancer. (1993) Some naturally occurring substances:

1222

food items and constituents, heterocyclic aromatic amines and mycotoxins. Lyon.

1223

(42) Waidyanatha, S., Lin, P. H., and Rappaport, S. M. (1996) Characterization of chlorinated

1224

adducts of hemoglobin and albumin following administration of pentachlorophenol to rats. Chem.

1225

Res. Toxicol. 9, 647–653.

1226

(43) Nguyen, T. N. T., Bertagnolli, A. D., Villalta, P. W., Bühlmann, P., and Sturla, S. J. (2005)

1227

Characterization

1228

tetrachlorobenzoquinone: dichlorobenzoquinone-1, N2-etheno-2‘-deoxyguanosine. Chem. Res.

1229

Toxicol. 18, 1770–1776.

1230

(44) Vaidyanathan, V. G., Villalta, P. W., and Sturla, S. J. (2007) Nucleobase-dependent

1231

reactivity of a quinone metabolite of pentachlorophenol. Chem. Res. Toxicol. 20, 913–919.

1232

(45) Dai, J., Park, G., Wright, M. W., Adams, M., Akman, S. A., and Manderville, R. A. (2002)

1233

Detection and characterization of a glutathione conjugate of ochratoxin A. Chem. Res. Toxicol.

1234

15, 1581–1588.

1235

(46) Gillman, I. G., Clark, T. N., and Manderville, R. A. (1999) Oxidation of ochratoxin A by an

1236

Fe-porphyrin system: model for enzymatic activation and DNA cleavage. Chem. Res. Toxicol.

1237

12, 1066–1076.

1238

(47) Hattemer-Frey, H. A., and Travis, C. C. (1989) Pentachlorophenol: environmental

1239

partitioning and human exposure. Arch. Environ. Contam. Toxicol. 18, 482–489.

of

a

deoxyguanosine

50

ACS Paragon Plus Environment

adduct

of

Page 51 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1240

(48) Jorens, P. G., and Schepens, P. J. (1993) Human pentachlorophenol poisoning. Hum. Exp.

1241

Toxicol. 12, 479–495.

1242

(49) Mcconnell, E. (1991) Toxicology and carcinogenesis studies of two grades of

1243

pentachlorophenol in B6C3F1 mice. Fund. Appl. Toxicol. 17, 519–532.

1244

(50) (1999) NTP toxicology and carcinogenesis studies of pentachlorophenol (CAS No. 87-86-5)

1245

in F344/n rats (feed studies). Natl. Toxicol. Program 483, 1-182.

1246

(51) Greene, M. H., Brinton, L. A., Fraumeni, J. F., and D'Amico, R. (1978) Familial and

1247

sporadic Hodgkin's disease associated with occupational wood exposure. Lancet 2, 626–627.

1248

(52) Hardell, L., and Sandström, A. (1979) Case-control study: soft-tissue sarcomas and exposure

1249

to phenoxyacetic acids or chlorophenols. Br. J. Cancer 39, 711–717.

1250

(53) Pearce, N. E., Smith, A. H., Howard, J. K., Sheppard, R. A., Giles, H. J., and Teague, C. A.

1251

(1986) Non-Hodgkin's lymphoma and exposure to phenoxyherbicides, chlorophenols, fencing

1252

work, and meat works employment: a case-control study. Br. J. Ind. Med. 43, 75–83.

1253

(54) Roberts, H. J. (1990) Pentachlorophenol-associated aplastic anemia, red cell aplasia,

1254

leukemia and other blood disorders. J. Fla. Med. Assoc. 77, 86–90.

1255

(55) US Department of Health and Human Services. (2001) Toxicological profile for

1256

pentachlorophenol. Agency for Toxic Substance and Disease Registry.

1257

(56) Renner, G., and Hopfer, C. (1990) Metabolic studies on pentachlorophenol (PCP) in rats.

1258

Xenobiotica 20, 573–582.

1259

(57) Proudfoot, A. T. (2003) Pentachlorophenol poisoning. Toxicol. Rev. 22, 3–11.

1260

(58) van Ommen, B., Voncken, J. W., Müller, F., and van Bladeren, P. J. (1988) The oxidation of

1261

tetrachloro-1,4-hydroquinone by microsomes and purified cytochrome P-450b. Implications for

1262

covalent binding to protein and involvement of reactive oxygen species. Chem. Biol. Interact. 65,

1263

247–259. 51

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 52 of 84

1264

(59) Lin, P. H., Waidyanatha, S., Pollack, G. M., Swenberg, J. A., and Rappaport, S. M. (1999)

1265

Dose-specific production of chlorinated quinone and semiquinone adducts in rodent livers

1266

following administration of pentachlorophenol. Toxicol. Sci. 47, 126–133.

1267

(60)

1268

tetrachlorohydroquinone on cell growth and the induction of DNA damage in Chinese hamster

1269

ovary cells. Mutat. Res. 244, 299–302.

1270

(61) Witte, I., Juhl, U., and Butte, W. (1985) DNA-damaging properties and cytotoxicity in

1271

human fibroblasts of tetrachlorohydroquinone, a pentachlorophenol metabolite. Mutat. Res. 145,

1272

71–75.

1273

(62) Zhu, B. Z., Kitrossky, N., and Chevion, M. (2000) Evidence for production of hydroxyl

1274

radicals by pentachlorophenol metabolites and hydrogen peroxide: a metal-independent organic

1275

Fenton reaction. Biochem. Biophys. Res. Commun. 270, 942–946.

1276

(63) Ahlborg, U. G., Larsson, K., and Thunberg, T. (1978) Metabolism of pentachlorophenol in

1277

vivo and in vitro. Arch. Toxicol. 40, 45–53.

1278

(64) Kang, P., Dalvie, D., Smith, E., and Renner, M. (2009) Bioactivation of lumiracoxib by

1279

peroxidases and human liver microsomes: identification of multiple quinone imine intermediates

1280

and GSH adducts. Chem. Res. Toxicol. 22, 106–117.

1281

(65) Li, Y., Slatter, J. G., Zhang, Z., Li, Y., Doss, G. A., Braun, M. P., Stearns, R. A., Dean, D.

1282

C., Baillie, T. A., and Tang, W. (2008) In vitro metabolic activation of lumiracoxib in rat and

1283

human liver preparations. Drug Metab. Dispos. 36, 469–473.

1284

(66) Poon, G. K., Chen, Q., Teffera, Y., Ngui, J. S., Griffin, P. R., Braun, M. P., Doss, G. A.,

1285

Freeden, C., Stearns, R. A., Evans, D. C., Baillie, T. A., and Tang, W. (2001) Bioactivation of

1286

diclofenac via benzoquinone imine intermediates-identification of urinary mercapturic acid

1287

derivatives in rats and humans. Drug Metab. Dispos. 29, 1608–1613.

Ehrlich,

W.

(1990)

The

effect

of

52

pentachlorophenol

ACS Paragon Plus Environment

and

its

metabolite

Page 53 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1288

(67) Shen, S., Marchick, M. R., Davis, M. R., Doss, G. A., and Pohl, L. R. (1999) Metabolic

1289

activation of diclofenac by human cytochrome P450 3A4: role of 5-hydroxydiclofenac. Chem.

1290

Res. Toxicol. 12, 214–222.

1291

(68) Tang, W., Stearns, R. A., Bandiera, S. M., Zhang, Y., Raab, C., Braun, M. P., Dean, D. C.,

1292

Pang, J., Leung, K. H., Doss, G. A., Strauss, J. R., Kwei, G. Y., Rushmore, T. H., Chiu, S. H.,

1293

and Baillie, T. A. (1999) Studies on cytochrome P-450-mediated bioactivation of diclofenac in

1294

rats and in human hepatocytes: identification of glutathione conjugated metabolites. Drug Metab.

1295

Dispos. 27, 365–372.

1296

(69) Smith, K. S., Smith, P. L., Heady, T. N., Trugman, J. M., Harman, W. D., and Macdonald,

1297

T. L. (2003) In vitro metabolism of tolcapone to reactive intermediates: relevance to tolcapone

1298

liver toxicity. Chem. Res. Toxicol. 16, 123–128.

1299

(70) Klopčič, I., Poberžnik, M., Mavri, J., and Dolenc, M. S. (2015) A quantum chemical study

1300

of the reactivity of acetaminophen (paracetamol) toxic metabolite N-acetyl-p-benzoquinone

1301

imine with deoxyguanosine and glutathione. Chem. Biol. Interact. 242, 407–414.

1302

(71) Neftel, K. A., Woodtly, W., Schmid, M., Frick, P. G., and Fehr, J. (1986) Amodiaquine

1303

induced agranulocytosis and liver damage. Br. Med. J. (Clin. Res. Ed.) 292, 721–723.

1304

(72) Maggs, J. L., Tingle, M. D., Kitteringham, N. R., and Park, B. K. (1988) Drug-protein

1305

conjugates-XIV. Mechanisms of formation of protein-arylating intermediates from amodiaquine,

1306

a myelotoxin and hepatotoxin in man. Biochem. Pharmacol. 37, 303–311.

1307

(73) Christie, G., Breckenridge, A. M., and Park, B. K. (1989) Drug-protein conjugates-XVIII.

1308

Detection of antibodies towards the antimalarial amodiaquine and its quinone imine metabolite in

1309

man and the rat. Biochem. Pharmacol. 38, 1451–1458.

1310

(74) Salama, A., and Mueller-Eckhardt, C. (1985) The role of metabolite-specific antibodies in

1311

nomifensine-dependent immune hemolytic anemia. N. Engl. J. Med. 313, 469–474. 53

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1312

(75) Yu, J., Mathisen, D. E., Burdette, D., Brown, D. G., Becker, C., and Aharony, D. (2010)

1313

Identification of multiple glutathione conjugates of 8-amino-2-methyl-4-phenyl-1,2,3,4-

1314

tetrahydroisoquinoline maleate (nomifensine) in liver microsomes and hepatocyte preparations:

1315

evidence of the bioactivation of nomifensine. Drug Metab. Dispos. 38, 46–60.

1316

(76) Teng, W. C., Oh, J. W., New, L. S., Wahlin, M. D., Nelson, S. D., Ho, H. K., and Chan, E.

1317

C. Y. (2010) Mechanism-based inactivation of cytochrome P450 3A4 by lapatinib. Mol.

1318

Pharmacol. 78, 693–703.

1319

(77) Morgan, R. E., Trauner, M., van Staden, C. J., Lee, P. H., Ramachandran, B., Eschenberg,

1320

M., Afshari, C. A., Qualls, C. W., Lightfoot-Dunn, R., and Hamadeh, H. K. (2010) Interference

1321

with bile salt export pump function is a susceptibility factor for human liver injury in drug

1322

development. Toxicol. Sci. 118, 485–500.

1323

(78) Li, X., He, Y., Ruiz, C. H., Koenig, M., Cameron, M. D., and Vojkovsky, T. (2009)

1324

Characterization of dasatinib and its structural analogs as CYP3A4 mechanism-based inactivators

1325

and the proposed bioactivation pathways. Drug Metab. Dispos. 37, 1242–1250.

1326

(79) Li, X., Kamenecka, T. M., and Cameron, M. D. (2009) Bioactivation of the epidermal

1327

growth factor receptor inhibitor gefitinib: implications for pulmonary and hepatic toxicities.

1328

Chem. Res. Toxicol. 22, 1736–1742.

1329

(80) Li, X., Kamenecka, T. M., and Cameron, M. D. (2010) Cytochrome P450-mediated

1330

bioactivation of the epidermal growth factor receptor inhibitor erlotinib to a reactive electrophile.

1331

Drug Metab. Dispos. 38, 1238–1245.

1332

(81) Kalgutkar, A. S., Vaz, A. D. N., Lame, M. E., Henne, K. R., Soglia, J., Zhao, S. X.,

1333

Abramov, Y. A., Lombardo, F., Collin, C., Hendsch, Z. S., and Hop, C. E. C. A. (2005)

1334

Bioactivation of the nontricyclic antidepressant nefazodone to a reactive quinone-imine species in

54

ACS Paragon Plus Environment

Page 54 of 84

Page 55 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1335

human liver microsomes and recombinant cytochrome P450 3A4. Drug Metab. Dispos. 33, 243–

1336

253.

1337

(82) Bauman, J. N., Frederick, K. S., Sawant, A., Walsky, R. L., Cox, L. M., Obach, R. S., and

1338

Kalgutkar, A. S. (2008) Comparison of the bioactivation potential of the antidepressant and

1339

hepatotoxin nefazodone with aripiprazole, a structural analog and marketed drug. Drug Metab.

1340

Dispos. 36, 1016–1029.

1341

(83) Greene, D. S., and Barbhaiya, R. H. (1997) Clinical Pharmacokinetics of Nefazodone. Clin.

1342

Pharmacokinet. 33, 260–275.

1343

(84) Kostrubsky, S. E., Strom, S. C., Kalgutkar, A. S., Kulkarni, S., Atherton, J., Mireles, R.,

1344

Feng, B., Kubik, R., Hanson, J., Urda, E., and Mutlib, A. E. (2006) Inhibition of hepatobiliary

1345

transport as a predictive method for clinical hepatotoxicity of nefazodone. Toxicol. Sci. 90, 451–

1346

459.

1347

(85) Dykens, J. A., Jamieson, J. D., Marroquin, L. D., Nadanaciva, S., Xu, J. J., Dunn, M. C.,

1348

Smith, A. R., and Will, Y. (2008) In vitro assessment of mitochondrial dysfunction and

1349

cytotoxicity of nefazodone, trazodone, and buspirone. Toxicol. Sci. 103, 335–345.

1350

(86) Shear, N. H., and Spielberg, S. P. (1988) Anticonvulsant hypersensitivity syndrome. In vitro

1351

assessment of risk. J. Clin. Invest. 82, 1826–1832.

1352

(87) Bu, H.-Z., Zhao, P., Dalvie, D. K., and Pool, W. F. (2007) Identification of primary and

1353

sequential bioactivation pathways of carbamazepine in human liver microsomes using liquid

1354

chromatography/tandem mass spectrometry. Rapid Commun. Mass Spectrom. 21, 3317–3322.

1355

(88) Bu, H.-Z., Kang, P., Deese, A. J., Zhao, P., and Pool, W. F. (2005) Human in vitro

1356

glutathionyl

1357

pharmacologically active metabolite of carbamazepine. Drug Metab. Dispos. 33, 1920–1924.

and

protein

adducts

of

carbamazepine-10,11-epoxide,

55

ACS Paragon Plus Environment

a

stable

and

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 56 of 84

1358

(89) Ju, C., and Uetrecht, J. P. (1999) Detection of 2-hydroxyiminostilbene in the urine of

1359

patients taking carbamazepine and its oxidation to a reactive iminoquinone intermediate. J.

1360

Pharmacol. Exp. Ther. 288, 51–56.

1361

(90) Lillibridge, J. H., Amore, B. M., Slattery, J. T., Kalhorn, T. F., Nelson, S. D., Finnell, R. H.,

1362

and Bennett, G. D. (1996) Protein-reactive metabolites of carbamazepine in mouse liver

1363

microsomes. Drug Metab. Dispos. 24, 509–514.

1364

(91) Nakayama, S., Atsumi, R., Takakusa, H., Kobayashi, Y., Kurihara, A., Nagai, Y., Nakai, D.,

1365

and Okazaki, O. (2009) A zone classification system for risk assessment of idiosyncratic drug

1366

toxicity using daily dose and covalent binding. Drug Metab. Dispos. 37, 1970–1977.

1367

(92) Lennernäs, H. (2003) Clinical pharmacokinetics of atorvastatin. Clin. Pharmacokinet. 42,

1368

1141–1160.

1369

(93) Kalgutkar, A. S., Henne, K. R., Lame, M. E., Vaz, A. D. N., Collin, C., Soglia, J. R., Zhao,

1370

S. X., and Hop, C. E. C. A. (2005) Metabolic activation of the nontricyclic antidepressant

1371

trazodone to electrophilic quinone-imine and epoxide intermediates in human liver microsomes

1372

and recombinant P4503A4. Chem. Biol. Interact. 155, 10–20.

1373

(94) Wen, B., Ma, L., Rodrigues, A. D., and Zhu, M. (2008) Detection of novel reactive

1374

metabolites

1375

chlorophenylpiperazine. Drug Metab. Dispos. 36, 841–850.

1376

(95) Jauch, R., Kopitar, Z., Prox, A., and Zimmer, A. (1976) Pharmacokinetics and metabolism

1377

of trazodone in man (author's transl). Arzneimittelforschung 26, 2084–2089.

1378

(96) Venkatakrishnan, K., and Obach, R. S. (2005) In vitro-in vivo extrapolation of CYP2D6

1379

inactivation by paroxetine: prediction of nonstationary pharmacokinetics and drug interaction

1380

magnitude. Drug Metab. Dispos. 33, 845–852.

of

trazodone:

evidence

for

56

CYP2D6-mediated

ACS Paragon Plus Environment

bioactivation

of

m-

Page 57 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1381

(97) Zhao, S. X., Dalvie, D. K., Kelly, J. M., Soglia, J. R., Frederick, K. S., Smith, E. B., Obach,

1382

R. S., and Kalgutkar, A. S. (2007) NADPH-dependent covalent binding of [3H]paroxetine to

1383

human liver microsomes and S-9 fractions: identification of an electrophilic quinone metabolite

1384

of paroxetine. Chem. Res. Toxicol. 20, 1649–1657.

1385

(98) Ring, B. J., Patterson, B. E., Mitchell, M. I., Vandenbranden, M., Gillespie, J., Bedding, A.

1386

W., Jewell, H., Payne, C. D., Forgue, S. T., Eckstein, J., Wrighton, S. A., and Phillips, D. L.

1387

(2005) Effect of tadalafil on cytochrome P450 3A4-mediated clearance: studies in vitro and in

1388

vivo. Clin. Pharmacol. Ther. 77, 63–75.

1389

(99) Lim, H.-K., Chen, J., Sensenhauser, C., Cook, K., and Subrahmanyam, V. (2007) Metabolite

1390

identification by data-dependent accurate mass spectrometric analysis at resolving power of

1391

60,000 in external calibration mode using an LTQ/Orbitrap. Rapid Commun. Mass Spectrom. 21,

1392

1821–1832.

1393

(100) Dalvie, D., Kang, P., Zientek, M., Xiang, C., Zhou, S., and Obach, R. S. (2008) Effect of

1394

intestinal glucuronidation in limiting hepatic exposure and bioactivation of raloxifene in humans

1395

and rats. Chem. Res. Toxicol. 21, 2260–2271.

1396

(101) Yu, L., Liu, H., Li, W., Zhang, F., Luckie, C., van Breemen, R. B., Thatcher, G. R. J., and

1397

Bolton, J. L. (2004) Oxidation of raloxifene to quinoids: potential toxic pathways via a diquinone

1398

methide and o-quinones. Chem. Res. Toxicol. 17, 879–888.

1399

(102) Li, K.-M., Todorovic, R., Devanesan, P., Higginbotham, S., Köfeler, H., Ramanathan, R.,

1400

Gross, M. L., Rogan, E. G., and Cavalieri, E. L. (2004) Metabolism and DNA binding studies of

1401

4-hydroxyestradiol and estradiol-3,4-quinone in vitro and in female ACI rat mammary gland in

1402

vivo. Carcinogenesis 25, 289–297.

57

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1403

(103) Stack, D. E., Byun, J., Gross, M. L., Rogan, E. G., and Cavalieri, E. L. (1996) Molecular

1404

characteristics of catechol estrogen quinones in reactions with deoxyribonucleosides. Chem. Res.

1405

Toxicol. 9, 851–859.

1406

(104) Bindhumol, V., Chitra, K. C., and Mathur, P. P. (2003) Bisphenol A induces reactive

1407

oxygen species generation in the liver of male rats. Toxicology 188, 117–124.

1408

(105) Ben-Jonathan, N., and Steinmetz, R. (1998) Xenoestrogens: the emerging story of

1409

bisphenol A. Trends Endocrin. Met. 9, 124–128.

1410

(106) Atkinson, A., and Roy, D. (1995) In vitro conversion of environmental estrogenic chemical

1411

bisphenol A to DNA binding metabolite(s). Biochem. Biophys. Res. Commun. 210, 424–433.

1412

(107) Kabuto, H., Hasuike, S., Minagawa, N., and Shishibori, T. (2003) Effects of bisphenol A

1413

on the metabolisms of active oxygen species in mouse tissues. Environ. Res. 93, 31–35.

1414

(108) Atkinson, A., and Roy, D. (1995) In vivo DNA adduct formation by bisphenol A. Environ.

1415

Mol. Mutagen. 26, 60–66.

1416

(109) Sajiki, J. (2001) Decomposition of bisphenol A (BPA) by radical oxygen. Environ. Int. 27,

1417

315–320.

1418

(110) Nicolis, S., Zucchelli, M., Monzani, E., and Casella, L. (2008) Myoglobin modification by

1419

enzyme-generated dopamine reactive species. Chemistry 14, 8661–8673.

1420

(111) Butler, J., and Hoey, B. M. (1987) Are reduced quinones necessarily involved in the

1421

antitumour activity of quinone drugs? Br. J. Cancer Suppl. 8, 53–59.

1422

(112) Minotti, G., Licata, S., Saponiero, A., Menna, P., Calafiore, A. M., Di Giammarco, G.,

1423

Liberi, G., Animati, F., Cipollone, A., Manzini, S., and Maggi, C. A. (2000) Anthracycline

1424

metabolism and toxicity in human myocardium: comparisons between doxorubicin, epirubicin,

1425

and a novel disaccharide analogue with a reduced level of formation and [4Fe-4S] reactivity of its

1426

secondary alcohol metabolite. Chem. Res. Toxicol. 13, 1336–1341. 58

ACS Paragon Plus Environment

Page 58 of 84

Page 59 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1427

(113) Maltoni, C., Ciliberti, A., Cotti, G., Conti, B., and Belpoggi, F. (1989) Benzene, an

1428

experimental multipotential carcinogen: results of the long-term bioassays performed at the

1429

Bologna Institute of Oncology. Environ. Health Perspect. 82, 109–124.

1430

(114) Maltoni, C., Conti, B., Cotti, G., and Belpoggi, F. (1985) Experimental studies on benzene

1431

carcinogenicity at the Bologna Institute of Oncology: current results and ongoing research. Am. J.

1432

Ind. Med. 7, 415–446.

1433

(115) Huff, J. E., Haseman, J. K., DeMarini, D. M., Eustis, S., Maronpot, R. R., Peters, A. C.,

1434

Persing, R. L., Chrisp, C. E., and Jacobs, A. C. (1989) Multiple-site carcinogenicity of benzene in

1435

Fischer 344 rats and B6C3F1 mice. Environ. Health Perspect. 82, 125–163.

1436

(116) Cronkite, E. P., Drew, R. T., Inoue, T., Hirabayashi, Y., and Bullis, J. E. (1989)

1437

Hematotoxicity and carcinogenicity of inhaled benzene. Environ. Health Perspect. 82, 97–108.

1438

(117) Merletti, F., Heseltine, E., Saracci, R., Simonato, L., Vainio, H., and Wilbourn, J. (1984).

1439

Target organs for carcinogenicity of chemicals and industrial exposures in humans: a review of

1440

results in the IARC monographs on the evaluation of the carcinogenic risk of chemicals to

1441

humans. Cancer Res. 44, 2244-2250.

1442

(118) Rinsky, R. A., Young, R. J., and Smith, A. B. (1981) Leukemia in benzene workers. Am. J.

1443

Ind. Med. 2, 217–245.

1444

(119) Ott, W. R., and Roberts, J. W. (1998) Everyday exposure to toxic pollutants. Sci. Am. 278,

1445

86–91.

1446

(120) Snyder, R., and Hedli, C. C. (1996) An overview of benzene metabolism. Environ. Health

1447

Perspect. 104, 1165–1171.

1448

(121) Sammett, D., Lee, E. W., Kocsis, J. J., and Snyder, R. (1979) Partial hepatectomy reduces

1449

both metabolism and toxicity of benzene. J. Toxicol. Environ. Health 5, 785–792.

59

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1450

(122) Andrews, L. S., Eun Woo Lee, Witmer, C. M., Kocsis, J. J., and Snyder, R. (1977) Effects

1451

of toluene on the metabolism, disposition and hemopoietic toxicity of [3H]benzene. Biochem.

1452

Pharmacol. 26, 293–300.

1453

(123) Seaton, M. J., Schlosser, P. M., Bond, J. A., and Medinsky, M. A. (1994) Benzene

1454

metabolism by human liver microsomes in relation to cytochrome P450 2E1 activity.

1455

Carcinogenesis 15, 1799–1806.

1456

(124) Valentine, J. L., Lee, S. S., Seaton, M. J., Asgharian, B., Farris, G., Corton, J. C., Gonzalez,

1457

F. J., and Medinsky, M. A. (1996) Reduction of benzene metabolism and toxicity in mice that

1458

lack CYP2E1 expression. Toxicol. Appl. Pharm. 141, 205–213.

1459

(125) Witz, G., Zhang, Z., and Goldstein, B. D. (1996) Reactive ring-opened aldehyde

1460

metabolites in benzene hematotoxicity. Environ. Health Perspect. 104, 1195–1199.

1461

(126) Uzma, N., Kumar, B., Salar, K., Madhuri, A., and Reddy, A. (2008) In vitro and in vivo

1462

evaluation of toxic effect of benzene on lymphocytes and hepatocytes. The Internet Journal of

1463

Toxicology. 6.

1464

(127) Tunek, A., Platt, K. L., Przybylski, M., and Oesch, F. (1980) Multi-step metabolic

1465

activation of benzene. Effect of superoxide dismutase on covalent binding to microsomal

1466

macromolecules, and identification of glutathione conjugates using high pressure liquid

1467

chromatography and field desorption mass spectrometry. Chem. Biol. Interact. 33, 1–17.

1468

(128) Rickert, D. E., Baker, T. S., Bus, J. S., Barrow, C. S., and Irons, R. D. (1979) Benzene

1469

disposition in the rat after exposure by inhalation. Toxicol. Appl. Pharm. 49, 417–423.

1470

(129) Greenlee, W. F., Gross, E. A., and Irons, R. D. (1981) Relationship between benzene

1471

toxicity and the disposition of 14C-labelled benzene metabolites in the rat. Chem. Biol. Interact.

1472

33, 285–299.

60

ACS Paragon Plus Environment

Page 60 of 84

Page 61 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1473

(130) Smart, R. C., and Zannoni, V. G. (1984) DT-diaphorase and peroxidase influence the

1474

covalent binding of the metabolites of phenol, the major metabolite of benzene. Mol. Pharmacol.

1475

26, 105–111.

1476

(131) Schlosser, M. J., and Kalf, G. F. (1989) Metabolic activation of hydroquinone by

1477

macrophage peroxidase. Chem. Biol. Interact. 72, 191–207.

1478

(132) Smith, M. T., Yager, J. W., Steinmetz, K. L., and Eastmond, D. A. (1989) Peroxidase-

1479

dependent metabolism of benzene's phenolic metabolites and its potential role in benzene toxicity

1480

and carcinogenicity. Environ. Health Perspect. 82, 23–29.

1481

(133) Twerdok, L. E., Rembish, S. J., and Trush, M. A. (1992) Induction of quinone reductase

1482

and glutathione in bone marrow cells by 1,2-dithiole-3-thione: effect on hydroquinone-induced

1483

cytotoxicity. Toxicol. Appl. Pharm. 112, 273–281.

1484

(134) Apel, K., and Hirt, H. (2004) Reactive oxygen species: metabolism, oxidative stress, and

1485

signal transduction. Annu. Rev. Plant Biol. 55, 373–399.

1486

(135) Guengerich, F. P., and Strickland, T. W. (1977) Metabolism of vinyl chloride: destruction

1487

of the heme of highly purified liver microsomal cytochrome P-450 by a metabolite. Mol.

1488

Pharmacol. 13, 993–1004.

1489

(136) Levin, W., Lu, A. Y. H., Jacobson, M., Kuntzman, R., Lee Poyer, J., and McCay, P. B.

1490

(1973) Lipid peroxidation and the degradation of cytochrome P-450 heme. Arch. Biochem.

1491

Biophys. 158, 842–852.

1492

(137) Schaefer, W. H., Harris, T. M., and Guengerich, F. P. (1985) Characterization of the

1493

enzymic and nonenzymic peroxidative degradation of iron porphyrins and cytochrome P-450

1494

heme. Biochemistry 24, 3254–3263.

1495

(138) Smith, M. T. (1996) The mechanism of benzene-induced leukemia: a hypothesis and

1496

speculations on the causes of leukemia. Environ. Health Perspect. 104, 1219–1225. 61

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1497

(139) Li, Y., Lafuente, A., and Trush, M. A. (1994) Characterization of quinone reductase,

1498

glutathione and glutathione S-transferase in human myeloid cell lines: induction by 1,2-dithiole-

1499

3-thione and effects on hydroquinone-induced cytotoxicity. Life Sci. 54, 901–916.

1500

(140) Ross, D., Siegel, D., Schattenberg, D. G., Sun, X. M., and Moran, J. L. (1996) Cell-specific

1501

activation and detoxification of benzene metabolites in mouse and human bone marrow:

1502

identification of target cells and a potential role for modulation of apoptosis in benzene toxicity.

1503

Environ. Health Perspect. 104, 1177–1182.

1504

(141) Li, Y., and Trush, M. A. (1993) DNA damage resulting from the oxidation of

1505

hydroquinone by copper: role for a Cu(II)/Cu(I) redox cycle and reactive oxygen generation.

1506

Carcinogenesis 14, 1303–1311.

1507

(142) Li, Y., Kuppusamy, P., Zweier, J. L., and Trush, M. A. (1995) ESR evidence for the

1508

generation of reactive oxygen species from the copper-mediated oxidation of the benzene

1509

metabolite, hydroquinone: role in DNA damage. Chem. Biol. Interact. 94, 101–120.

1510

(143) Li, Y., Kuppusamy, P., Zweier, J. L., and Trush, M. A. (1996) Role of Cu/Zn-superoxide

1511

dismutase in xenobiotic activation. I. Chemical reactions involved in the Cu/Zn-superoxide

1512

dismutase-accelerated oxidation of the benzene metabolite 1,4-hydroquinone. Mol. Pharmacol.

1513

49, 404–411.

1514

(144) Li, Y., Kuppusamy, P., Zweir, J. L., and Trush, M. A. (1996) Role of Cu/Zn-superoxide

1515

dismutase in xenobiotic activation. II. Biological effects resulting from the Cu/Zn-superoxide

1516

dismutase-accelerated oxidation of the benzene metabolite 1,4-hydroquinone. Mol. Pharmacol.

1517

49, 412–421.

1518

(145) Rothman, N., Smith, M. T., Hayes, R. B., Traver, R. D., Hoener, B., Campleman, S., Li, G.

1519

L., Dosemeci, M., Linet, M., Zhang, L., Xi, L., Wacholder, S., Lu, W., Meyer, K. B., Titenko-

1520

Holland, N., Stewart, J. T., Yin, S., and Ross, D. (1997) Benzene poisoning, a risk factor for 62

ACS Paragon Plus Environment

Page 62 of 84

Page 63 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

609

1521

hematological malignancy, is associated with the NQO1

1522

excretion of chlorzoxazone. Cancer Res. 57, 2839–2842.

1523

(146) Moran, J. L., Siegel, D., and Ross, D. (1999) A potential mechanism underlying the

1524

increased susceptibility of individuals with a polymorphism in NAD(P)H:quinone oxidoreductase

1525

1 (NQO1) to benzene toxicity. Proc. Natl. Acad. Sci. U.S.A. 96, 8150–8155.

1526

(147) Trush, M. A., Twerdok, L. E., Rembish, S. J., Zhu, H., and Li, Y. (1996) Analysis of target

1527

cell susceptibility as a basis for the development of a chemoprotective strategy against benzene-

1528

induced hematotoxicities. Environ. Health Perspect. 104, 1227–1234.

1529

(148) Lee, E. W., Kocsis, J. J., and Snyder, R. (1981) The use of ferrokinetics in the study of

1530

experimental anemia. Environ. Health Perspect. 39, 29–37.

1531

(149) Orzechowski, A., Schwarz, L. R., Schwegler, U., Bock, K. W., Snyder, R., and Schrenk, D.

1532

(1995) Benzene metabolism in rodent hepatocytes: role of sulphate conjugation. Xenobiotica 25,

1533

1093–1102.

1534

(150) Sanctucci, K., and Shah, B. (2000) Association of naphthalene with acute hemolytic

1535

anemia. Acad. Emergency Med. 7, 42–47.

1536

(151) Shi, H., Sui, Y., Wang, X., Luo, Y., and Ji, L. (2005) Hydroxyl radical production and

1537

oxidative damage induced by cadmium and naphthalene in liver of Carassius auratus. Comp.

1538

Biochem. Physiol. C Toxicol. Pharmacol. 140, 115–121.

1539

(152) Collén, J., Pinto, E., Pedersén, M., and Colepicolo, P. (2003) Induction of oxidative stress

1540

in the red macroalga Gracilaria tenuistipitata by pollutant metals. Arch. Environ. Contam.

1541

Toxicol. 45, 337–342.

1542

(153) Vijayavel, K., Gomathi, R. D., Durgabhavani, K., and Balasubramanian, M. P. (2004)

1543

Sublethal effect of naphthalene on lipid peroxidation and antioxidant status in the edible marine

1544

crab Scylla serrata. Mar. Pollut. Bull. 48, 429–433. 63

C-->T mutation and rapid fractional

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1545

(154) Pacheco, M., and Santos, M. A. (2002) Naphthalene and β-naphthoflavone effects on

1546

Anguilla anguilla L. hepatic metabolism and erythrocytic nuclear abnormalities. Environ. Int. 28,

1547

285–293.

1548

(155) Levitan, W. M., and Taylor, M. H. (1979) Physiology of salinity-dependent naphthalene

1549

toxicity in Fundulus heteroclitus. J. Fish. Res. Bd. Can. 36, 615–620.

1550

(156) Teles, M., Pacheco, M., and Santos, M. A. (2003) Anguilla anguilla L. liver

1551

ethoxyresorufin O-deethylation, glutathione S-tranferase, erythrocytic nuclear abnormalities, and

1552

endocrine responses to naphthalene and β-naphthoflavone. Ecotoxicol. Environ. Saf. 55, 98–107.

1553

(157) Preuss, R., Angerer, J. R., and Drexler, H. (2003) Naphthalene-an environmental and

1554

occupational toxicant. Int. Arch. Occup. Environ. Health 76, 556–576.

1555

(158) Simmonds, M. S. J. (2004) IARC monographs on the evaluation of carcinogenic risks to

1556

humans. Vol. 82, some traditional herbal medicines, some mycotoxins, naphthalene and styrene.

1557

Phytochemistry 65, 139.

1558

(159) Cho, T. M., Rose, R. L., and Hodgson, E. (2006) In vitro metabolism of naphthalene by

1559

human liver microsomal cytochrome P450 enzymes. Drug Metab. Dispos. 34, 176–183.

1560

(160) Stohs, S. (2002) Naphthalene toxicity and antioxidant nutrients. Toxicology 180, 97–105.

1561

(161) Jerina, D. M., Daly, J. W., Witkop, B., Zaltzman-Nirenberg, P., and Udenfriend, S. (1970)

1562

1,2-Naphthalene oxide as an intermediate in the microsomal hydroxylation of naphthalene.

1563

Biochemistry 9, 147–156.

1564

(162) Tsuruda, L. S., Lamé, M. W., and Jones, A. D. (1995) Formation of epoxide and quinone

1565

protein adducts in B6C3F1 mice treated with naphthalene, sulfate conjugate of 1,4-

1566

dihydroxynaphthalene and 1,4-naphthoquinone. Arch. Toxicol. 69, 362–367.

64

ACS Paragon Plus Environment

Page 64 of 84

Page 65 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1567

(163) Bakke, J., Struble, C., Gustafsson, J. A., and Gustafsson, B. (1985) Catabolism of

1568

premercapturic acid pathway metabolites of naphthalene to naphthols and methylthio-containing

1569

metabolites in rats. Proc. Natl. Acad. Sci. U.S.A. 82, 668–671.

1570

(164) Boyland, E., and Sims, P. (1958) Metabolism of polycyclic compounds. 12. An acid-labile

1571

precursor

1572

dihydrohydroxynaphthyl)-l-cysteine. Biochem. J. 68, 440–447.

1573

(165) Chen, K. C., and Dorough, H. W. (1979) Glutathione and mercapturic acid conjugations in

1574

the metabolism of naphthalene and 1-naphthyl N-methylcarbamate (carbaryl). Drug Chem.

1575

Toxicol. 2, 331–354.

1576

(166) Pakenham, G., Lango, J., Buonarati, M., Morin, D., and Buckpitt, A. (2002) Urinary

1577

naphthalene mercapturates as biomarkers of exposure and stereoselectivity of naphthalene

1578

epoxidation. Drug Metab. Dispos. 30, 247–253.

1579

(167) Rozman, K., Summer, K. H., Rozman, T., and Greim, H. (1982) Elimination of thioethers

1580

following administration of naphthalene and diethylmaleate to the rhesus monkey. Drug Chem.

1581

Toxicol. 5, 265–275.

1582

(168) Stillwell, W. G., Bouwsma, O. J., Thenot, J. P., Horning, M. G., Griffin, G. W., Ishikawa,

1583

K., and Takaku, M. (1978) Methylthio metabolites of naphthalen excreted by the rat. Res.

1584

Commun. Chem. Pathol. Pharmacol. 20, 509–530.

1585

(169) Yu, D., Berlin, J. A., Penning, T. M., and Field, J. (2002) Reactive oxygen species

1586

generated by PAH o-quinones cause change-in-function mutations in p53. Chem. Res. Toxicol.

1587

15, 832–842.

1588

(170) Wilson, A. S., Davis, C. D., Williams, D. P., Buckpitt, A. R., Pirmohamed, M., and Park,

1589

B. K. (1996) Characterisation of the toxic metabolite(s) of naphthalene. Toxicology 114, 233–

1590

242.

of

1-naphthylmercapturic

acid

65

and

naphthol:

ACS Paragon Plus Environment

an

N-acetyl-S-(1:2-

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1591

(171) Bagchi, D., Balmoori, J., Bagchi, M., Ye, X., Williams, C. B., and Stohs, S. J. (2000) Role

1592

of p53 tumor suppressor gene in the toxicity of TCDD, endrin, naphthalene, and chromium (vi) in

1593

liver and brain tissues of mice. Free Radic. Biol. Med. 28, 895–903.

1594

(172) Denissenko, M. F., Pao, A., Tang, M., and Pfeifer, G. P. (1996) Preferential formation of

1595

benzo[a]pyrene adducts at lung cancer mutational hotspots in p53. Science 274, 430–432.

1596

(173) Greene, J. F., Zheng, J., Grant, D. F., and Hammock, B. D. (2000) Cytotoxicity of 1,2-

1597

epoxynaphthalene is correlated with protein binding and in situ glutathione depletion in

1598

cytochrome P4501A1 expressing Sf-21 cells. Toxicol. Sci. 53, 352–360.

1599

(174) Hemminki, K. (1993) DNA adducts, mutations and cancer. Carcinogenesis 14, 2007–2012.

1600

(175) Troester, M. A., Lindstrom, A. B., Waidyanatha, S., Kupper, L. L., and Rappaport, S. M.

1601

(2002) Stability of hemoglobin and albumin adducts of naphthalene oxide, 1,2-naphthoquinone,

1602

and 1,4-naphthoquinone. Toxicol. Sci. 68, 314–321.

1603

(176) Waidyanatha, S., Troester, M. A., Lindstrom, A. B., and Rappaport, S. M. (2002)

1604

Measurement of hemoglobin and albumin adducts of naphthalene-1,2-oxide, 1,2-naphthoquinone

1605

and 1,4-naphthoquinone after administration of naphthalene to F344 rats. Chem. Biol. Interact.

1606

141, 189–210.

1607

(177) O'Driscoll, C., Doonan, F., Sanvicens, N., Messeguer, A., and Cotter, T. G. (2011) A novel

1608

free radical scavenger rescues retinal cells in vivo. Exp. Eye Res. 93, 65–74.

1609

(178) Yenes, S., and Messeguer, A. (1999) A study of the reaction of different phenol substrates

1610

with nitric oxide and peroxynitrite. Tetrahedron 55, 14111–14122.

1611

(179) Pérez-Asensio, F. J., la Rosa, de, X., Jiménez-Altayó, F., Gorina, R., Martínez, E.,

1612

Messeguer, A., Vila, E., Chamorro, A., and Planas, A. M. (2010) Antioxidant CR-6 protects

1613

against reperfusion injury after a transient episode of focal brain ischemia in rats. J. Cereb. Blood

1614

Flow Metab. 30, 638–652. 66

ACS Paragon Plus Environment

Page 66 of 84

Page 67 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1615

(180) Jiménez-Altayó, F., Caracuel, L., Pérez-Asensio, F. J., Martínez-Revelles, S., Messeguer,

1616

A., Planas, A. M., and Vila, E. (2009) Participation of oxidative stress on rat middle cerebral

1617

artery changes induced by focal cerebral ischemia: beneficial effects of 3,4-dihydro-6-hydroxy-7-

1618

methoxy-2,2-dimethyl-1(2H)-benzopyran (CR-6). J. Pharmacol. Exp. Ther. 331, 429–436.

1619

(181) Montoliu, C., Llansola, M., Sáez, R., Yenes, S., Messeguer, A., and Felipo, V. (1999)

1620

Prevention of glutamate neurotoxicity in cultured neurons by 3,4-dihydro-6-hydroxy-7-methoxy-

1621

2,2-dimethyl-1(2H)-benzopyran (CR-6), a scavenger of nitric oxide. Biochem. Pharmacol. 58,

1622

255–261.

1623

(182) Miranda, M., Muriach, M., Almansa, I., Arnal, E., Messeguer, A., Díaz-Llopis, M.,

1624

Romero, F. J., and Bosch-Morell, F. (2007) CR-6 protects glutathione peroxidase activity in

1625

experimental diabetes. Free Radic. Biol. Med. 43, 1494–1498.

1626

(183) Sanvicens, N., Gómez-Vicente, V., Masip, I., Messeguer, A., and Cotter, T. G. (2004)

1627

Oxidative stress-induced apoptosis in retinal photoreceptor cells is mediated by calpains and

1628

caspases and blocked by the oxygen radical scavenger CR-6. J. Biol. Chem. 279, 39268–39278.

1629

(184) Sanvicens, N., Gómez-Vicente, V., Messeguer, A., and Cotter, T. G. (2006) The radical

1630

scavenger CR-6 protects SH-SY5Y neuroblastoma cells from oxidative stress-induced apoptosis:

1631

effect on survival pathways. J. Neurochem. 98, 735–747.

1632

(185) Morgenstern, R., DePierre, J. W., and Jörnvall, H. (1985) Microsomal glutathione

1633

transferase. Primary structure. J. Biol. Chem. 260, 13976–13983.

1634

(186) Commandeur, J. N., Stijntjes, G. J., and Vermeulen, N. P. (1995) Enzymes and transport

1635

systems involved in the formation and disposition of glutathione S-conjugates. Role in

1636

bioactivation and detoxication mechanisms of xenobiotics. Pharmacol. Rev. 47, 271–330.

67

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1637

(187) Morgenstern, R., and DePierre, J. W. (1983) Microsomal glutathione transferase.

1638

Purification in unactivated form and further characterization of the activation process, substrate

1639

specificity and amino acid composition. Eur. J. Biochem. 134, 591–597.

1640

(188) Andersson, C., Weinander, R., Lundqvist, G., DePierre, J. W., and Morgenstern, R. (1994)

1641

Functional and structural membrane topology of rat liver microsomal glutathione transferase.

1642

Biochim. Biophys. Acta 1204, 298–304.

1643

(189) van der Merwe, K. J., Steyn, P. S., and Fourie, L. (1965) Ochratoxin A, a toxic metabolite

1644

produced by Aspergillus ochraceus Wilh. Nature 205, 1112–1113.

1645

(190) Pfohl-Leszkowicz, A., and Manderville, R. A. (2007) Ochratoxin A: an overview on

1646

toxicity and carcinogenicity in animals and humans. Mol. Nutr. Food Res. 51, 61–99.

1647

(191) World Health Organization. (2001) Evaluation of certain mycotoxins in food. Fifty-sixth

1648

report of the joint FAO/WHO expert committee on food additives. World Health Organ Tech

1649

Rep Ser. 906, 1-62.

1650

(192) Manderville, R. A. (2005) A case for the genotoxicity of ochratoxin A by bioactivation and

1651

covalent DNA adduction. Chem. Res. Toxicol. 18, 1091–1097.

1652

(193) Turesky, R. J. (2005) Perspective: ochratoxin A is not a genotoxic carcinogen. Chem. Res.

1653

Toxicol. 18, 1082–1090.

1654

(194) Castegnaro, M. (1991) Mycotoxins, endemic nephropathy and urinary tract tumours (Vol.

1655

115). WHO.

1656

(195) Petzinger, E., and Ziegler, K. (2000) Ochratoxin A from a toxicological perspective. J. Vet.

1657

Pharmacol. Ther. 23, 91–98.

1658

(196) Chu, F. S., Noh, I., and Chang, C. C. (1972) Structural requirements for ochratoxin

1659

intoxication. Life Sci. 11, 503–508.

68

ACS Paragon Plus Environment

Page 68 of 84

Page 69 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1660

(197) Bredenkamp, M. W., Dillen, J. L. M., van Rooyen, P. H., and Steyn, P. S. (1989) Crystal

1661

structures and conformational analysis of ochratoxin A and B: probing the chemical structure

1662

causing toxicity. J. Chem. Soc. 2, 1835–1839.

1663

(198) Föllmann, W., Hillebrand, I. E., Creppy, E. E., and Bolt, H. M. (1995) Sister chromatid

1664

exchange frequency in cultured isolated porcine urinary bladder epithelial cells (PUBEC) treated

1665

with ochratoxin A and alpha. Arch. Toxicol. 69, 280–286.

1666

(199) Marin-Kuan, M., Ehrlich, V., Delatour, T., Cavin, C., and Schilter, B. (2011) Evidence for

1667

a role of oxidative stress in the carcinogenicity of ochratoxin A. J. Toxicol. 2011, 645361–15.

1668

(200) Gautier, J. C., Holzhaeuser, D., Markovic, J., Gremaud, E., Schilter, B., and Turesky, R. J.

1669

(2001) Oxidative damage and stress response from ochratoxin A exposure in rats. Free Radic.

1670

Biol. Med. 30, 1089–1098.

1671

(201) Schaaf, G. J., Nijmeijer, S. M., Maas, R. F. M., Roestenberg, P., de Groene, E. M., and

1672

Fink-Gremmels, J. (2002) The role of oxidative stress in the ochratoxin A-mediated toxicity in

1673

proximal tubular cells. Biochim. Biophys. Acta 1588, 149–158.

1674

(202) Kamp, H. G., Eisenbrand, G., Janzowski, C., Kiossev, J., Latendresse, J. R., Schlatter, J.,

1675

and Turesky, R. J. (2005) Ochratoxin A induces oxidative DNA damage in liver and kidney after

1676

oral dosing to rats. Mol. Nutr. Food Res. 49, 1160–1167.

1677

(203) Mally, A., Pepe, G., Ravoori, S., Fiore, M., Gupta, R. C., Dekant, W., and Mosesso, P.

1678

(2005) Ochratoxin A causes DNA damage and cytogenetic effects but no DNA adducts in rats.

1679

Chem. Res. Toxicol. 18, 1253–1261.

1680

(204) Kamp, H. G., Eisenbrand, G., Schlatter, J., Würth, K., and Janzowski, C. (2005)

1681

Ochratoxin A: induction of (oxidative) DNA damage, cytotoxicity and apoptosis in mammalian

1682

cell lines and primary cells. Toxicology 206, 413–425.

69

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1683

(205) Arbillaga, L., Azqueta, A., Ezpeleta, O., and López de Cerain, A. (2007) Oxidative DNA

1684

damage induced by ochratoxin A in the HK-2 human kidney cell line: evidence of the

1685

relationship with cytotoxicity. Mutagenesis 22, 35–42.

1686

(206) Ali, R., Mittelstaedt, R. A., Shaddock, J. G., Ding, W., Bhalli, J. A., Khan, Q. M., and

1687

Heflich, R. H. (2011) Comparative analysis of micronuclei and DNA damage induced by

1688

ochratoxin A in two mammalian cell lines. Mutat. Res. 723, 58–64.

1689

(207) Shigenaga, M. K., Gimeno, C. J., and Ames, B. N. (1989) Urinary 8-hydroxy-2'-

1690

deoxyguanosine as a biological marker of in vivo oxidative DNA damage. Proc. Natl. Acad. Sci.

1691

U.S.A. 86, 9697–9701.

1692

(208) Cavin, C., Delatour, T., Marin-Kuan, M., Fenaille, F., Holzhäuser, D., Guignard, G.,

1693

Bezençon, C., Piguet, D., Parisod, V., Richoz-Payot, J., and Schilter, B. (2009) Ochratoxin A-

1694

mediated DNA and protein damage: roles of nitrosative and oxidative stresses. Toxicol. Sci. 110,

1695

84–94.

1696

(209) Murray, A. R., Kisin, E., Castranova, V., Kommineni, C., Gunther, M. R., and Shvedova,

1697

A. A. (2007) Phenol-induced in vivo oxidative stress in skin: evidence for enhanced free radical

1698

generation, thiol oxidation, and antioxidant depletion. Chem. Res. Toxicol. 20, 1769–1777.

1699

(210) Stoyanovosky, D. A., Goldman, R., Jonnalagadda, S. S., Day, B. W., Claycamp, H. G., and

1700

Kagan, V. E. (1996) Detection and characterization of the electron paramagnetic resonance-silent

1701

glutathionyl-5,5-dimethyl-1-pyrroline N-oxide adduct derived from redox cycling of phenoxyl

1702

radicals in model systems and HL-60 cells. Arch. Biochem. Biophys. 330, 3–11.

1703

(211) Mally, A., and Dekant, W. (2009) Mycotoxins and the kidney: modes of action for renal

1704

tumor formation by ochratoxin A in rodents. Mol. Nutr. Food Res. 53, 467–478.

1705

(212) Rached, E., Pfeiffer, E., Dekant, W., and Mally, A. (2006) Ochratoxin A: apoptosis and

1706

aberrant exit from mitosis due to perturbation of microtubule dynamics? Toxicol. Sci. 92, 78–86. 70

ACS Paragon Plus Environment

Page 70 of 84

Page 71 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1707

(213) Adler, M., Müller, K., Rached, E., Dekant, W., and Mally, A. (2009) Modulation of key

1708

regulators of mitosis linked to chromosomal instability is an early event in ochratoxin A

1709

carcinogenicity. Carcinogenesis 30, 711–719.

1710

(214) Czakai, K., Müller, K., Mosesso, P., Pepe, G., Schulze, M., Gohla, A., Patnaik, D., Dekant,

1711

W., Higgins, J. M. G., and Mally, A. (2011) Perturbation of mitosis through inhibition of histone

1712

acetyltransferases: the key to ochratoxin a toxicity and carcinogenicity? Toxicol. Sci. 122, 317–

1713

329.

1714

(215) Bhandari, S., Kalowski, S., Collett, P., Cooke, B. E., Kerr, P., Newland, R., Dowling, J.,

1715

and Horvath, J. (2002) Karyomegalic nephropathy: an uncommon cause of progressive renal

1716

failure. Nephrol. Dial. Transplant. 17, 1914–1920.

1717

(216) Pfohl-Leszkowicz, A., Bartsch, H., Azémar, B., Mohr, U., Estève, J., Castegnaro, M.

1718

(2002) MESNA protects rats against nephrotoxicity but not carcinogenicity induced by

1719

ochratoxin A, implicating two separate pathways. Facta Univ. Ser. Med. 9, 57-63.

1720

(217) Lin, P.-H., La, D. K., Upton, P. B., and Swenberg, J. A. (2002) Analysis of DNA adducts

1721

in rats exposed to pentachlorophenol. Carcinogenesis 23, 365–369.

1722

(218) Samokyszyn, V. M., Freeman, J. P., Maddipati, K. R., and Lloyd, R. V. (1995) Peroxidase-

1723

catalyzed oxidation of pentachlorophenol. Chem. Res. Toxicol. 8, 349–355.

1724

(219) Dai, J., Wright, M. W., and Manderville, R. A. (2003) An oxygen-bonded C8-

1725

deoxyguanosine nucleoside adduct of pentachlorophenol by peroxidase activation: evidence for

1726

ambident C8 reactivity by phenoxyl radicals. Chem. Res. Toxicol. 16, 817–821.

1727

(220) Dai, J., Sloat, A. L., Wright, M. W., and Manderville, R. A. (2005) Role of phenoxyl

1728

radicals in DNA adduction by chlorophenol xenobiotics following peroxidase activation. Chem.

1729

Res. Toxicol. 18, 771–779.

71

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 72 of 84

1730

(221) Lawson, T., Gannett, P. M., Yau, W.-M., Dalal, N. S., and Toth, B. (1995) Different

1731

patterns of mutagenicity of arenediazonium ions in V79 cells and Salmonella typhimurium

1732

TA102: evidence for different mechanisms of action. J. Agric. Food Chem. 43, 2627–2635.

1733

(222) Hiramoto, K., Kaku, M., Sueyoshi, A., Fujise, M., and Kikugawa, K. (1995) DNA base and

1734

deoxyribose

1735

(hydroxymethyl)benzenediazonium salt, a carcinogen in mushroom. Chem. Res. Toxicol. 8, 356–

1736

362.

1737

(223) Gannett, P. M., Powell, J. H., Rao, R., Shi, X., Lawson, T., Kolar, C., and Toth, B. (1999)

1738

C8-Arylguanine and C8-aryladenine formation in calf thymus DNA from arenediazonium ions.

1739

Chem. Res. Toxicol. 12, 297–304.

1740

(224) Dai, J., Wright, M. W., and Manderville, R. A. (2003) Ochratoxin A forms a carbon-

1741

bonded C8-deoxyguanosine nucleoside adduct: implications for C8 reactivity by a phenolic

1742

radical. J. Am. Chem. Soc. 125, 3716–3717.

1743

(225) Il'ichev, Y. V., Perry, J. L., Manderville, R. A., Chignell, C. F., and Simon, J. D. (2001)

1744

The pH-dependent primary photoreactions of ochratoxin A. J. Phys. Chem. B 105, 11369–11376.

1745

(226) Manderville, R. A. (2009) Structural and biological impact of radical addition reactions

1746

with DNA nucleobases. Adv. Phys. Org. Chem. 43, 177–218.

1747

(227) Faucet, V., Pfohl-Leszkowicz, A., Dai, J., Castegnaro, M., and Manderville, R. A. (2004)

1748

Evidence for covalent DNA adduction by ochratoxin A following chronic exposure to rat and

1749

subacute exposure to pig. Chem. Res. Toxicol. 17, 1289–1296.

1750

(228) Manderville, R., and Pfohl-Leszkowicz, A. (2008) Bioactivation and DNA adduction as a

1751

rationale for ochratoxin A carcinogenesis. WMJ 1, 357–367.

modification

by

the

carbon-centered

72

ACS Paragon Plus Environment

radical

generated

from

4-

Page 73 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1752

(229) Umemura, T., Sai-Kato, K., Takagi, A., Hasegawa, R., and Kurokawa, Y. (1996) Oxidative

1753

DNA damage and cell proliferation in the livers of B6C3F1 mice exposed to pentachlorophenol

1754

in their diet. Fundam. Appl. Toxicol. 30, 285–289.

1755

(230) Morgan, D. P. (1989) Recognition and management of pesticide poisonings. DIANE

1756

Publishing.

1757

(231) Colosio, C., Maroni, M., Barcellini, W., Meroni, P., Alcini, D., Colombi, A., Cavallo, D.,

1758

and Foa, V. (1993) Toxicological and immune findings in workers exposed to pentachlorophenol

1759

(PCP). Arch. Environ. Health 48, 81–88.

1760

(232) Wilson, N. K., Chuang, J. C., Morgan, M. K., Lordo, R. A., and Sheldon, L. S. (2007) An

1761

observational study of the potential exposures of preschool children to pentachlorophenol,

1762

bisphenol A, and nonylphenol at home and daycare. Environ. Res. 103, 9–20.

1763

(233) Thompson, T. S., and Treble, R. G. (1994) Preliminary results of a survey of

1764

pentachlorophenol levels in human urine. Bull. Environ. Contam. Toxicol. 53, 274–279.

1765

(234) Carrizo, D., Grimalt, J. O., Ribas-Fito, N., Torrent, M., and Sunyer, J. (2008)

1766

Pentachlorobenzene, hexachlorobenzene, and pentachlorophenol in children's serum from

1767

industrial and rural populations after restricted use. Ecotoxicol. Environ. Saf. 71, 260–266.

1768

(235) Daniel, V., Huber, W., Bauer, K., Suesal, C., Mytilineos, J., Melk, A., Conradt, C., and

1769

Opelz, G. (2001) Association of elevated blood levels of pentachlorophenol (PCP) with cellular

1770

and humoral immunodeficiencies. Arch. Environ. Health 56, 77–83.

1771

(236) Uhl, S., Schmid, P., and Schlatter, C. (1986) Pharmacokinetics of pentachlorophenol in

1772

man. Arch. Toxicol. 58, 182–186.

1773

(237) Triebig, G., Krekeler, H., Gossler, K. and Valentin, H. (1980) Investigations on

1774

neurotoxicity of chemical substances at the workplace. II. Determination of the motor and

73

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1775

sensory nerve conduction velocity in persons occupationally exposed to pentachlorophenol

1776

(author's transl). Int. Arch. Occup. Environ. Health 48, 357–367.

1777

(238) Wang, Y. J., Ho, Y. S., Jeng, J. H., Su, H. J., and Lee, C. C. (2000) Different cell death

1778

mechanisms and gene expression in human cells induced by pentachlorophenol and its major

1779

metabolite, tetrachlorohydroquinone. Chem. Biol. Interact. 128, 173–188.

1780

(239) O'Donoghue, J. L. (1985) Neurotoxicity of industrial and commercial chemicals. Boca

1781

Raton, Fla, CRC Press.

1782

(240) Dorsey, W. C., Tchounwou, P. B., and Ford, B. D. (2006) Neuregulin 1-beta cytoprotective

1783

role in AML 12 mouse hepatocytes exposed to pentachlorophenol. IJERPH 3, 11–22.

1784

(241) Lin, P. H., Waidyanatha, S., Pollack, G. M., and Rappaport, S. M. (1997) Dosimetry of

1785

chlorinated quinone metabolites of pentachlorophenol in the livers of rats and mice based upon

1786

measurement of protein adducts. Toxicol. Appl. Pharm. 145, 399–408.

1787

(242) Tsai, C.-H., Lin, P.-H., Waidyanatha, S., and Rappaport, S. M. (2001) Characterization of

1788

metabolic activation of pentachlorophenol to quinones and semiquinones in rodent liver. Chem.

1789

Biol. Interact. 134, 55–71.

1790

(243) Besten, den, C., van Bladeren, P. J., Duizer, E., Vervoort, J., and Rietjens, I. M. (1993)

1791

Cytochrome P450-mediated oxidation of pentafluorophenol to tetrafluorobenzoquinone as the

1792

primary reaction product. Chem. Res. Toxicol. 6, 674–680.

1793

(244) Rietjens, I. M., Besten, den, C., Hanzlik, R. P., and van Bladeren, P. J. (1997) Cytochrome

1794

P450-catalyzed oxidation of halobenzene derivatives. Chem. Res. Toxicol. 10, 629–635.

1795

(245) Lin, P. H., Waidyanatha, S., and Rappaport, S. M. (1996) Investigation of liver binding of

1796

pentachlorophenol based upon measurement of protein adducts. Biomarkers 1, 232–243.

74

ACS Paragon Plus Environment

Page 74 of 84

Page 75 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1797

(246) Mangold, J. B., Gu, H., Rodriguez, L. C., Bonner, J., Dickson, J., and Rordorf, C. (2004)

1798

Pharmacokinetics and metabolism of lumiracoxib in healthy male subjects. Drug Metab. Dispos.

1799

32, 566–571.

1800

(247) Jorga, K., Fotteler, B., Heizmann, P., and Gasser, R. (1999) Metabolism and excretion of

1801

tolcapone, a novel inhibitor of catechol-O-methyltransferase. Br. J. Clin. Pharmacol. 48, 513–

1802

520.

1803

(248) Martignoni, E., Cosentino, M., Ferrari, M., Porta, G., Mattarucchi, E., Marino, F., Lecchini,

1804

S., and Nappi, G. (2005) Two patients with COMT inhibitor-induced hepatic dysfunction and

1805

UGT1A9 genetic polymorphism. Neurology 65, 1820–1822.

1806

(249) Korlipara, L. V. P., Cooper, J. M., and Schapira, A. H. V. (2004) Differences in toxicity of

1807

the catechol-O-methyl transferase inhibitors, tolcapone and entacapone to cultured human

1808

neuroblastoma cells. Neuropharmacology 46, 562–569.

1809

(250) Jensen, M. S., Rebordosa, C., Thulstrup, A. M., Toft, G., Sørensen, H. T., Bonde, J. P.,

1810

Henriksen, T. B., and Olsen, J. (2010) Maternal use of acetaminophen, ibuprofen, and

1811

acetylsalicylic acid during pregnancy and risk of cryptorchidism. Epidemiology 21, 779–785.

1812

(251) Kristensen, D. M., Hass, U., Lesné, L., Lottrup, G., Jacobsen, P. R., Desdoits-Lethimonier,

1813

C., Boberg, J., Petersen, J. H., Toppari, J., Jensen, T. K., Brunak, S., Skakkebaek, N. E.,

1814

Nellemann, C., Main, K. M., Jégou, B., and Leffers, H. (2011) Intrauterine exposure to mild

1815

analgesics is a risk factor for development of male reproductive disorders in human and rat. Hum.

1816

Reprod. 26, 235–244.

1817

(252) Isidori, M., Bellotta, M., Cangiano, M., and Parrella, A. (2009) Estrogenic activity of

1818

pharmaceuticals in the aquatic environment. Environ. Int. 35, 826–829.

1819

(253) Fent, K., Escher, C., and Caminada, D. (2006) Estrogenic activity of pharmaceuticals and

1820

pharmaceutical mixtures in a yeast reporter gene system. Reprod. Toxicol. 22, 175–185. 75

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1821

(254) Kristensen, D. M., Lesné, L., Le Fol, V., Desdoits-Lethimonier, C., Dejucq-Rainsford, N.,

1822

Leffers, H., and Jégou, B. (2012) Paracetamol (acetaminophen), aspirin (acetylsalicylic acid) and

1823

indomethacin are anti-androgenic in the rat foetal testis. Int. J. Androl. 35, 377–384.

1824

(255) IARC Working Group on the Evaluation of Carcinogenic Risks to Humans and

1825

International Agency for Research on Cancer. (1999) Some chemicals that cause tumours of the

1826

kidney or urinary bladder in rodents and some other substances.

1827

(256) Court, M. H., Duan, S. X., Moltke, von, L. L., Greenblatt, D. J., Patten, C. J., Miners, J. O.,

1828

and Mackenzie, P. I. (2001) Interindividual variability in acetaminophen glucuronidation by

1829

human liver microsomes: identification of relevant acetaminophen UDP-glucuronosyltransferase

1830

isoforms. J. Pharmacol. Exp. Ther. 299, 998–1006.

1831

(257) Mutlib, A. E., Goosen, T. C., Bauman, J. N., Williams, J. A., Kulkarni, S., and Kostrubsky,

1832

S. (2006) Kinetics of acetaminophen glucuronidation by UDP-glucuronosyltransferases 1A1,

1833

1A6, 1A9 and 2B15. Potential implications in acetaminophen-induced hepatotoxicity. Chem. Res.

1834

Toxicol. 19, 701–709.

1835

(258) Patten, C. J., Thomas, P. E., Guy, R. L., Lee, M., Gonzalez, F. J., Guengerich, F. P., and

1836

Yang, C. S. (1993) Cytochrome P450 enzymes involved in acetaminophen activation by rat and

1837

human liver microsomes and their kinetics. Chem. Res. Toxicol. 6, 511–518.

1838

(259) Lee, S. S. T., Buters, J. T. M., Pineau, T., Fernandez-Salguero, P., and Gonzalez, F. J.

1839

(1996) Role of CYP2E1 in the hepatotoxicity of acetaminophen. J. Biol. Chem. 271, 12063–

1840

12067.

1841

(260) Stiborová, M., Mikšanová, M., Havlı́ček, V., Schmeiser, H. H., and Frei, E. (2002)

1842

Mechanism of peroxidase-mediated oxidation of carcinogenic o-anisidine and its binding to

1843

DNA. Mutat. Res. 500, 49–66.

76

ACS Paragon Plus Environment

Page 76 of 84

Page 77 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1844

(261) Tan, S., Sagara, Y., Liu, Y., Maher, P., and Schubert, D. (1998) The regulation of reactive

1845

oxygen species production during programmed cell death. J. Cell Biol. 141, 1423–1432.

1846

(262) Jeff Slepin, M. (1984) Acetaminophen overdose in young children. J. Emerg. Med. 2, 148–

1847

149.

1848

(263) Churchill, F. C., Patchen, L. C., Campbell, C. C., Schwartz, I. K., Nguyen-Dinh, P., and

1849

Dickinson, C. M. (1985) Amodiaquine as a prodrug: importance of metabolite(s) in the

1850

antimalarial effect of amodiaquine in humans. Life Sci. 36, 53–62.

1851

(264) O'Neill, P. M., Harrison, A. C., Storr, R. C., Hawley, S. R., Ward, S. A., and Park, B. K.

1852

(1994) The effect of fluorine substitution on the metabolism and antimalarial activity of

1853

amodiaquine. J. Med. Chem. 37, 1362–1370.

1854

(265) O'Neill, P. M., Shone, A. E., Stanford, D., Nixon, G., Asadollahy, E., Park, B. K., Maggs,

1855

J. L., Roberts, P., Stocks, P. A., Biagini, G., Bray, P. G., Davies, J., Berry, N., Hall, C., Rimmer,

1856

K., Winstanley, P. A., Hindley, S., Bambal, R. B., Davis, C. B., Bates, M., Gresham, S. L.,

1857

Brigandi, R. A., Gomez-de-Las-Heras, F. M., Gargallo, D. V., Parapini, S., Vivas, L., Lander, H.,

1858

Taramelli, D., and Ward, S. A. (2009) Synthesis, antimalarial activity, and preclinical

1859

pharmacology of a novel series of 4'-fluoro and 4'-chloro analogues of amodiaquine.

1860

Identification of a suitable “back-up” compound for N-tert-butyl isoquine. J. Med. Chem. 52,

1861

1828–1844.

1862

(266) Spraggs, C. F., Budde, L. R., Briley, L. P., Bing, N., Cox, C. J., King, K. S., Whittaker, J.

1863

C., Mooser, V. E., Preston, A. J., Stein, S. H., and Cardon, L. R. (2011) HLA-DQA1*02:01 is a

1864

major risk factor for lapatinib-induced hepatotoxicity in women with advanced breast cancer. J.

1865

Clin. Oncol. 29, 667–673.

1866

(267) Lertratanangkoon, K., and Horning, M. G. (1982) Metabolism of carbamazepine. Drug

1867

Metab. Dispos. 10, 1–10. 77

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 78 of 84

1868

(268) Maggs, J. L., Pirmohamed, M., Kitteringham, N. R., and Park, B. K. (1997)

1869

Characterization

1870

chromatography/mass spectrometry. Drug Metab. Dispos. 25, 275–280.

1871

(269) Pirmohamed, M., Kitteringham, N. R., Guenthner, T. M., Breckenridge, A. M., and Park,

1872

B. K. (1992) An investigation of the formation of cytotoxic, protein-reactive and stable

1873

metabolites from carbamazepine in vitro. Biochem. Pharmacol. 43, 1675–1682.

1874

(270) Spielberg, S. P., Gordon, G. B., Blake, D. A., Mellits, E. D., and Bross, D. S. (1981)

1875

Anticonvulsant toxicity in vitro: possible role of arene oxides. J. Pharmacol. Exp. Ther. 217,

1876

386–389.

1877

(271) Pearce, R. E., Lu, W., Wang, Y., Uetrecht, J. P., Correia, M. A., and Leeder, J. S. (2008)

1878

Pathways of carbamazepine bioactivation in vitro. III. The role of human cytochrome P450

1879

enzymes in the formation of 2,3-dihydroxycarbamazepine. Drug Metab. Dispos. 36, 1637–1649.

1880

(272) Leeder, J. S., Gaedigk, A., Lu, X., and Cook, V. A. (1996) Epitope mapping studies with

1881

human anti-cytochrome P450 3A antibodies. Mol. Pharmacol. 49, 234–243.

1882

(273) Paoletti, R., Corsini, A., and Bellosta, S. (2002) Pharmacological interactions of statins.

1883

Atherosclerosis Supp. 3, 35–40.

1884

(274) Fernandes, N. F., Martin, R. R., and Schenker, S. (2000) Trazodone-induced

1885

hepatotoxicity: a case report with comments on drug-induced hepatotoxicity. Am. J.

1886

Gastroenterol. 95, 532–535.

1887

(275) Zhou, H. H., and Wood, A. J. (1995) Stereoselective disposition of carvedilol is determined

1888

by CYP2D6. Clin. Pharmacol. Ther. 57, 518–524.

1889

(276) Kemp, D. C., Fan, P. W., and Stevens, J. C. (2002) Characterization of raloxifene

1890

glucuronidation in vitro: contribution of intestinal metabolism to presystemic clearance. Drug

1891

Metab. Dispos. 30, 694–700.

of

the

metabolites

of

carbamazepine

78

ACS Paragon Plus Environment

in

patient

urine

by

liquid

Page 79 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1892

(277) Russo, J., and Russo, I. H. (2006) The role of estrogen in the initiation of breast cancer. J.

1893

Steroid Biochem. Mol. Biol. 102, 89–96.

1894

(278) Cavalieri, E. L., and Rogan, E. G. (2011) Unbalanced metabolism of endogenous estrogens

1895

in the etiology and prevention of human cancer. J. Steroid Biochem. Mol. Biol. 125, 169–180.

1896

(279) Ptak, A., and Gregoraszczuk, E. L. (2013) Oestrogens, xenoestrogens and hormone-

1897

dependent cancers. In Carcinogenesis (Tonissen, K., Ed.) pp 63-86, InTech.

1898

(280) Kitamura, S., Suzuki, T., Sanoh, S., Kohta, R., Jinno, N., Sugihara, K., Yoshihara, S.,

1899

Fujimoto, N., Watanabe, H., and Ohta, S. (2005) Comparative study of the endocrine-disrupting

1900

activity of bisphenol A and 19 related compounds. Toxicol. Sci. 84, 249–259.

1901

(281) Xin, F., Jiang, L., Liu, X., Geng, C., Wang, W., Zhong, L., Yang, G., and Chen, M. (2014)

1902

Bisphenol A induces oxidative stress-associated DNA damage in INS-1 cells. Mutat. Res. Genet.

1903

Toxicol. Environ. Mutagen. 769, 29–33.

1904

(282) Brotons, J. A., Olea-Serrano, M. F., Villalobos, M., Pedraza, V., and Olea, N. (1995)

1905

Xenoestrogens released from lacquer coatings in food cans. Environ. Health Perspect. 103, 608–

1906

612.

1907

(283) Hashimoto, Y., Moriguchi, Y., Oshima, H., Kawaguchi, M., Miyazaki, K., and Nakamura,

1908

M. (2001) Measurement of estrogenic activity of chemicals for the development of new dental

1909

polymers. Toxicol. in Vitro 15, 421–425.

1910

(284) Olea, N., Pulgar, R., Pérez, P., Olea-Serrano, F., Rivas, A., Novillo-Fertrell, A., Pedraza,

1911

V., Soto, A. M., and Sonnenschein, C. (1996) Estrogenicity of resin-based composites and

1912

sealants used in dentistry. Environ. Health Perspect. 104, 298–305.

1913

(285) Kabuto, H., Amakawa, M., and Shishibori, T. (2004) Exposure to bisphenol A during

1914

embryonic/fetal life and infancy increases oxidative injury and causes underdevelopment of the

1915

brain and testis in mice. Life Sci. 74, 2931–2940. 79

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1916

(286) Yang, Y. J., Hong, Y.-C., Oh, S.-Y., Park, M.-S., Kim, H., Leem, J.-H., and Ha, E.-H.

1917

(2009) Bisphenol A exposure is associated with oxidative stress and inflammation in

1918

postmenopausal women. Environ. Res. 109, 797–801.

1919

(287) Chitra, K. (2003) Induction of oxidative stress by bisphenol A in the epididymal sperm of

1920

rats. Toxicology 185, 119–127.

1921

(288) Obata, T., Kinemuchi, H., and Aomine, M. (2002) Protective effect of diltiazem, a L-type

1922

calcium channel antagonist, on bisphenol A-enhanced hydroxyl radical generation by 1-methyl-4-

1923

phenylpyridinium ion in rat striatum. Neurosci. Lett. 334, 211–213.

1924

(289) Benachour, N., and Aris, A. (2009) Toxic effects of low doses of bisphenol A on human

1925

placental cells. Toxicol. Appl. Pharm. 241, 322–328.

1926

(290) Domoradzki, J. Y., Thornton, C. M., Pottenger, L. H., Hansen, S. C., Card, T. L.,

1927

Markham, D. A., Dryzga, M. D., Shiotsuka, R. N., and Waechter, J. M. (2004) Age and dose

1928

dependency of the pharmacokinetics and metabolism of bisphenol A in neonatal sprague-dawley

1929

rats following oral administration. Toxicol. Sci. 77, 230–242.

1930

(291) Snyder, R. W., Maness, S. C., Gaido, K. W., Welsch, F., Sumner, S. C., and Fennell, T. R.

1931

(2000) Metabolism and disposition of bisphenol A in female rats. Toxicol. Appl. Pharm. 168,

1932

225–234.

1933

(292) Völkel, W., Colnot, T., Csanády, G. A., Filser, J. G., and Dekant, W. (2002) Metabolism

1934

and kinetics of bisphenol A in humans at low doses following oral administration. Chem. Res.

1935

Toxicol. 15, 1281–1287.

1936

(293) Domoradzki, J. Y. (2003) Metabolism and pharmacokinetics of bisphenol A (BPA) and the

1937

embryo-fetal distribution of BPA and BPA-monoglucuronide in CD Sprague-Dawley rats at three

1938

gestational stages. Toxicol. Sci. 76, 21–34.

80

ACS Paragon Plus Environment

Page 80 of 84

Page 81 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1939

(294) Pritchett, J. J., Kuester, R. K., and Sipes, I. G. (2002) Metabolism of bisphenol A in

1940

primary cultured hepatocytes from mice, rats, and humans. Drug Metab. Dispos. 30, 1180–1185.

1941

(295) Hayashi, O., Kameshiro, M., Masuda, M., and Satoh, K. (2008) Bioaccumulation and

1942

metabolism of [14C]bisphenol A in the brackish water bivalve Corbicula japonica. Biosci.

1943

Biotechnol. Biochem. 72, 3219–3224.

1944

(296) Fini, J.-B., Dolo, L., Cravedi, J.-P., Demeneix, B., and Zalko, D. (2009) Metabolism of the

1945

endocrine disruptor BPA by Xenopus laevis Tadpoles. Ann. NY Acad. Sci. 1163, 394–397.

1946

(297) Yoshida, M., Ono, H., Mori, Y., Chuda, Y., and Mori, M. (2002) Oxygenation of bisphenol

1947

A to quinones by polyphenol oxidase in vegetables. J. Agric. Food Chem. 50, 4377–4381.

1948

(298) Yoshida, M., Ono, H., Mori, Y., Chuda, Y., and Onishi, K. (2001) Oxidation of bisphenol

1949

A and related compounds. Biosci. Biotechnol. Biochem. 65, 1444–1446.

1950

(299) Stokes, A. H., Hastings, T. G., and Vrana, K. E. (1999) Cytotoxic and genotoxic potential

1951

of dopamine. J. Neurosci. Res. 55, 659–665.

1952

(300) Halliwell, B. (2006) Reactive oxygen species and the central nervous system. J.

1953

Neurochem. 59, 1609–1623.

1954

(301) Donaldson, J., McGregor, D., and LaBella, F. (1982) Manganese neurotoxicity: a model for

1955

free radical mediated neurodegeneration? Can. J. Physiol. Pharmacol. 60, 1398–1405.

1956

(302) Halliwell, B., and Gutteridge, J. M. (1984) Oxygen toxicity, oxygen radicals, transition

1957

metals and disease. Biochem. J. 219, 1–14.

1958

(303) Wick, M., Byers, L., and Frei, E. (1977) L-dopa: selective toxicity for melanoma cells in

1959

vitro. Science 197, 468–469.

1960

(304) Graham, D. G. (1978) Oxidative pathways for catecholamines in the genesis of

1961

neuromelanin and cytotoxic quinones. Mol. Pharmacol. 14, 633–643.

81

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1962

(305) Hastings, T. G. (2002) Enzymatic oxidation of dopamine: the role of prostaglandin H

1963

synthase. J. Neurochem. 64, 919–924.

1964

(306) Napolitano, A., Crescenzi, O., Pezzella, A., and Prota, G. (1995) Generation of the

1965

neurotoxin 6-hydroxydopamine by peroxidase/H2O2 oxidation of dopamine. J. Med. Chem. 38,

1966

917–922.

1967

(307) Stokes, A. H., Brown, B. G., Lee, C. K., Doolittle, D. J., and Vrana, K. E. (1996)

1968

Tyrosinase enhances the covalent modification of DNA by dopamine. Brain Res. Mol. Brain Res.

1969

42, 167–170.

1970

(308) Weiss, R. B. (1992) The anthracyclines: will we ever find a better doxorubicin? Semin.

1971

Oncol. 19, 670–686.

1972

(309) Long, B. H., Musial, S. T., and Brattain, M. G. (1984) Comparison of cytotoxicity and

1973

DNA breakage activity of congeners of podophyllotoxin including VP16-213 and VM26: a

1974

quantitative structure-activity relationship. Biochemistry 23, 1183–1188.

1975

(310) Nitiss, J. L. (2009) Targeting DNA topoisomerase II in cancer chemotherapy. Nat. Rev.

1976

Cancer 9, 338–350.

1977

(311) Sinha, B. K., and Mason, R. P. (2015) Is metabolic activation of topoisomerase II poisons

1978

important in the mechanism of cytotoxicity? J. Drug Metab. Toxicol. 6, 186.

1979

(312) Bachur, N. R., Gordon, S. L., Gee, M. V., and Kon, H. (1979) NADPH cytochrome P-450

1980

reductase activation of quinone anticancer agents to free radicals. Proc. Natl. Acad. Sci. U.S.A.

1981

76, 954–957.

1982

(313) Sato, S., Iwaizumi, M., Handa, K., and Tamura, Y. (1977) Electron spin resonance study on

1983

the mode of generation of free radicals of daunomycin, adriamycin, and carboquone in

1984

NAD(P)H-microsome system. Gan 68, 603–608.

82

ACS Paragon Plus Environment

Page 82 of 84

Page 83 of 84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemical Research in Toxicology

1985

(314) Kalyanaraman, B., Nemec, J., and Sinha, B. K. (1989) Characterization of free radicals

1986

produced during oxidation of etoposide (VP-16) and its catechol and quinone derivatives. An

1987

ESR study. Biochemistry 28, 4839–4846.

1988

(315) Sinha, B. K., Katki, A. G., Batist, G., Cowan, K. H., and Myers, C. E. (1987) Adriamycin-

1989

stimulated hydroxyl radical formation in human breast tumor cells. Biochem. Pharmacol. 36,

1990

793–796.

1991

(316) Sinha, B. K., Katki, A. G., Batist, G., Cowan, K. H., and Myers, C. E. (1987) Differential

1992

formation of hydroxyl radicals by adriamycin in sensitive and resistant MCF-7 human breast

1993

tumor cells: implications for the mechanism of action. Biochemistry 26, 3776–3781.

1994

(317) Sinha, B. K., and Chignell, C. F. (1979) Binding mode of chemically activated

1995

semiquinone free radicals from quinone anticancer agents to DNA. Chem. Biol. Interact. 28,

1996

301–308.

1997

(318) Sinha, B. K. (1980) Binding specificity of chemically and enzymatically activated

1998

anthracycline anticancer agents to nucleic acids. Chem. Biol. Interact. 30, 67–77.

1999

(319) Sinha, B. K., and Sik, R. H. (1980) Binding of [14C]-adriamycin to cellular

2000

macromolecules in vivo. Biochem. Pharmacol. 29, 1867–1868.

2001

(320) Sinha, B. K., Trush, M. A., Kennedy, K. A., and Mimnaugh, E. G. (1984) Enzymatic

2002

activation and binding of adriamycin to nuclear DNA. Cancer Res. 44, 2892–2896.

2003

(321) Cullinane, C., and Phillips, D. R. (1990) Induction of stable transcriptional blockage sites

2004

by adriamycin: GpC specificity of apparent adriamycin-DNA adducts and dependence on

2005

iron(III) ions. Biochemistry 29, 5638–5646.

2006

(322) Cullinane, C., van Rosmalen, A., and Phillips, D. R. (1994) Does adriamycin induce

2007

interstrand cross-links in DNA? Biochemistry 33, 4632–4638.

83

ACS Paragon Plus Environment

Chemical Research in Toxicology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2008

(323) Cullinane, C. (2000) Interstrand cross-linking by adriamycin in nuclear and mitochondrial

2009

DNA of MCF-7 cells. Nucleic Acids Res. 28, 1019–1025.

2010

(324) Cutts, S., Nudelman, A., Rephaeli, A., and Phillips, D. (2005) The power and potential of

2011

doxorubicin-DNA adducts. IUBMB Life 57, 73–81.

2012

(325) van Rosmalen, A., Cullinane, C., Cutts, S. M., and Phillips, D. R. (1995) Stability of

2013

adriamycin-induced DNA adducts and interstrand crosslinks. Nucleic Acids Res. 23, 42–50.

2014 2015 2016 2017 2018 2019 2020 2021 2022 2023 2024 2025 2026 2027 2028 2029 2030 2031

Biographies: Marija Sollner Dolenc is Full Professor at the Faculty of Pharmacy, University of Ljubljana. She obtained her Ph.D. in the field of Medicinal Chemistry at the same faculty. Her work was first focused on design and synthesis of biologically active compounds (immunomodulators, antibacterial and anticancer agents). In the last few years, she works on the field of toxicology; in particular she is engaged in the research of formation and toxicity of reactive metabolites, while the second research area is evaluation of potential endocrine disrupting chemicals and their mixtures. Ivana Klopčič is a researcher scientist in Department of Dissolution Studies, Slovenian Development Center, Lek Pharmaceuticals, d.d. a Sandoz Company, where she focuses on validation of analytical procedures for dissolution test of new pharmaceutical drugs. She obtained M. Sc. degree in the study program Pharmacy at the Faculty of Pharmacy, University of Novi Sad, Serbia; and Ph.D. degree in the study program Toxicology at the Faculty of Pharmacy, University of Ljubljana, Slovenia. She has supervised one B.Sc. student and is author of seven publications in peer-reviewed journals. Her Ph. D. thesis focused on in-silico and in-vitro evaluation of potential endocrine disrupting chemicals and their mixtures.

84

ACS Paragon Plus Environment

Page 84 of 84