Cancer Letters

Cancer Letters

Volume 318, Issue 2, 28 May 2012, Pages 206-213
Cancer Letters

Reduced 4-aminobiphenyl-induced liver tumorigenicity but not DNA damage in arylamine N-acetyltransferase null mice

https://doi.org/10.1016/j.canlet.2011.12.022Get rights and content

Abstract

The aromatic amine 4-aminobiphenyl (ABP) is a liver procarcinogen in mice, requiring enzymatic bioactivation to exert its tumorigenic effect. To assess the role of arylamine N-acetyltransferase (NAT)-dependent acetylation capacity in the risk for ABP-induced liver tumors, we compared 1-year liver tumor incidence following the postnatal exposure of wild-type and NAT-deficient Nat1/2(−/−) mice to ABP. At an ABP exposure of 1200 nmol, male Nat1/2(−/−) mice had a liver tumor incidence of 36% compared to 69% in wild-type males, and at 600 nmol there was a complete absence of tumors compared to 60% in wild-type mice. Only one female wild-type mouse had a tumor using this exposure protocol. However, levels of N-deoxyguanosin-8-yl-ABP-DNA adducts did not correlate with either the strain or sex differences in tumor incidence. These results suggest that female sex and NAT deficiency reduce risk for ABP-induced liver tumors, but by mechanisms unrelated to differences in DNA-damaging events.

Introduction

Metabolic bioactivation of chemical carcinogens into DNA damaging agents is a causal factor in a multi-step process leading to the development of tissue-specific tumors [1]. While genotoxic effects are believed to be a necessary prerequisite for chemically-induced carcinogenesis, it has become clear that non-genotoxic effects also contribute to the development of tumors [2], [3] by sustaining or promoting a proliferative environment, which is one of the proposed hallmarks of cancer [4].

The aromatic amine 4-aminobiphenyl (ABP), which is found in cigarette smoke and hair dyes, is a recognized human bladder carcinogen (for review see [3]). Since environmental exposure to ABP may thus contribute to the occurrence of some tissue-specific cancers, many studies have been undertaken to measure and correlate potential biomarkers of ABP exposure and cancer risk. ABP-DNA adducts have been detected in the DNA from bladder cancer patients [5], exfoliated ductal epithelial cells isolated from human breast milk [6] and from hepatocytes obtained from hepatocellular carcinoma cases [7], and a higher urinary excretion of ABP is found in smokers versus non-smokers [8]. In addition, ABP can bind to proteins such as hemoglobin and albumin, and thus protein adducts such as ABP hemoglobin adducts have also been used as a surrogate or biomarker for exposure and potential risk [9].

The first step in one proposed bioactivation model for ABP is its N-oxidation to N-hydroxy-ABP (N-OH-ABP). The hepatic cytochrome P450 (CYP) isoform CYP1A2 has been implicated in this reaction in the liver [10], whereas CYP1A1 [11] or peroxidases [12] may be involved in its localized oxidation within other tissues. O-acetylation (OAT) of N-OH-ABP by arylamine N-acetyltransferases (NATs) can yield an acetoxy ester, which is chemically unstable and decomposes into a highly reactive arylnitrenium ion that can bind to and damage DNA. The major DNA adduct formed as a result of ABP bioactivation is N-deoxyguanosin-8-yl-ABP (dG-C8-ABP) which causes primarily G:C  T:A transversion mutations [13] that have been mapped as a common mutation in codon 61 of the H-ras gene of ABP-exposed mice [14]. Competing with N-oxidation in the liver is the protective NAT-mediated N-acetylation which leads to the production of an innocuous acetamide metabolite. Thus, it is unclear whether NAT enzymes facilitate (by O-acetylation) or prevent (by N-acetylation) ABP-induced tumorigenesis.

Two functional NAT isozymes, NAT1 and NAT2, are present in humans and display distinct tissue distributions and substrate specificities [15]. In human liver, both ABP N-acetylation and N-OH-ABP O-acetylation activities are mediated predominantly by NAT2 [16], whereas NAT1 has a widespread tissue distribution and may be involved in the localized activation of aromatic amines within other tissues such as the bladder [17]. In addition, NAT1 is transcribed by alternatively spliced promoters, NATa and NATb, giving rise to multiple transcripts which differ in tissue specificity and the ability to form ABP-DNA adducts and mutations in vitro[18]. Thus, the role that NAT enzymes may have in either the bioactivation or detoxification of aromatic amines such as ABP may ultimately depend on their tissue expression profiles, the balance of competing enzyme pathways, the inducibility of these pathways, the dose and route of exposure, and organ-specific metabolism that can have a significant effect on the rate of clearance [19], [20]. Furthermore, NAT enzyme activity can be modulated by environmental factors such as oxidative stress arising from inflammation or environmental exposure [21], and by folate levels which have been shown to affect the methylation status of the mouse Nat2 promoter [22]. Thus, genotype is not the only factor that can influence functional NAT enzyme activity.

Since human NAT1 and NAT2 are both highly polymorphic enzymes (for review see [15], [23]), several epidemiological studies have been undertaken to investigate the role of rapid versus slow acetylator phenotype or NAT genotype to tissue-specific cancer risk, particularly in relation to arylamine or heterocyclic amine exposure. NAT1 and/or NAT2 genotypes have been investigated with respect to cancers of the bladder, colon, breast and liver [24], [25], [26], [27]. Although results from epidemiological studies have tended to be contradictory and indicate that there is no clear or consistent correlation between acetylator status and cancer risk, more recent meta-analyses have indicated an increased risk of either breast or bladder cancer in NAT2 slow acetylators that is dependent on smoking intensity [25], [26]. Limitations to epidemiological studies include small sample sizes, genotype misclassification errors and lack of quantitative information on chemical exposures due to the complex mixtures of aromatic amines and other chemicals to which individuals may be exposed.

To investigate the role of NAT-dependent acetylation in ABP-induced tumorigenesis, in the present study we have used the neonatal carcinogen bioassay and our Nat1/2(−/−) mice, which are deficient in both mouse Nat1 and Nat2 enzyme activities [28]. The neonatal mouse carcinogen bioassay is an accelerated carcinogenicity assay that has been validated using ABP as a positive control for liver tumors [29]. Using this assay, chemical carcinogens requiring metabolic bioactivation tend to show a restricted tissue tumor sensitivity that is reflective of the distribution of available drug metabolizing enzymes, whereas direct-acting genotoxic agents such as the alkylating agent N-ethyl-N-nitrosourea are more likely to result in a widespread tissue distribution of tumors. On the other hand, non-genotoxic chemical carcinogens that mediate increased lipid peroxidation or oxidative stress do not produce tumors using this bioassay [30]. The assay has also been used to demonstrate that in contrast to the bioactivation model described above, mouse Cyp1a2 may not be the initial enzyme involved in ABP-induced carcinogenesis [31]. ABP N-oxidation activity was still evident in Cyp1a2(−/−) mice, albeit at reduced levels, and ABP-DNA adduct levels were not reduced in these animals [32]. These findings have all challenged the view that mouse Cyp1a2 is involved in the metabolic bioactivation of aromatic amines such as ABP, and reinforces the necessity for conducting bioactivation and carcinogenicity studies invivo to determine the role that specific biotransformation enzymes such as the NATs may play in determining tissue-specific tumor risk.

Section snippets

Materials

All animal handling procedures were conducted in accordance with Canadian Council on Animal Care guidelines, and were approved by the University of Toronto Animal Care Committee. ABP, carnitine acetyltransferase, acetyl-DL-carnitine and acetyl-CoA were purchased from Sigma–Aldrich Canada Ltd (Oakville, ON, Canada). Enzymes used for digesting genomic DNA into monomers (DNase I, micrococcal nuclease, nuclease P1, spleen phosphodiesterase, snake venom phosphodiesterase, and alkaline phosphatase)

Tumor incidence and multiplicity

There was no significant difference in mean body weights between wild-type and Nat1/2(−/−) mice of the same treatment group, with the exception of male Nat1/2(−/−) mice treated with the 600 nmol dose of ABP (Table 1). These mice weighed less than the corresponding wild-type male mice treated with 600 nmol ABP (p < 0.05). Liver weights were not significantly different between mice of the same treatment group except for male wild-type versus male Nat1/2(−/−) mice and female wild-type versus female

Discussion

We have used the neonatal mouse carcinogen bioassay and our Nat1- and Nat2-deficient Nat1/2(−/−) mice to interrogate the role of the NAT enzymes in either facilitating or preventing ABP-induced liver tumorigenesis. In our study, ABP-induced liver tumorigenesis was either attenuated (at the higher ABP dose) or completely absent (at the lower ABP dose) in male Nat1/2(−/−) mice compared to male wild-type mice at the ABP doses used. On the surface, these results would suggest that in the absence of

Acknowledgments

We thank Christopher Koo, Aldora Ho and Daniel Hanna for their technical assistance. This research was supported by operating grants from the Canadian Institutes of Health Research and the National Cancer Institute of Canada (DMG), and by USPHS Grant [R01-CA034627] from the National Cancer Institute (DWH).

References (57)

  • J.M. Tiang et al.

    Small molecule inhibition of arylamine N-acetyltransferase Type I inhibits proliferation and invasiveness of MDA-MB-231 breast cancer cells

    Biochem. Biophys. Res. Commun.

    (2010)
  • A. Kawamura et al.

    Mouse N-acetyltransferase type 2, the homologue of human N-acetyltransferase type 1

    Biochem. Pharmacol.

    (2008)
  • W. Cao et al.

    N-acetyltransferase 2 activity and folate levels

    Life Sci.

    (2010)
  • A. Ward et al.

    Arylamine N-acetyltransferase in human red blood cells

    Biochem. Pharmacol.

    (1992)
  • D.C. Cabelof et al.

    Imbalanced base excision repair in response to folate deficiency is accelerated by polymerase beta haploinsufficiency

    J. Biol. Chem.

    (2004)
  • H.G. Neumann

    Aromatic amines in experimental cancer research: tissue-specific effects, an old problem and new solutions

    Crit. Rev. Toxicol.

    (2007)
  • S.M. Cohen et al.

    4-Aminobiphenyl and DNA reactivity: case study within the context of the 2006 IPCS Human Relevance Framework for Analysis of a cancer mode of action for humans

    Crit. Rev. Toxicol.

    (2006)
  • B. Zayas et al.

    Detection and quantification of 4-ABP adducts in DNA from bladder cancer patients

    Carcinogenesis

    (2007)
  • L.Y. Wang et al.

    4-Aminobiphenyl DNA damage in liver tissue of hepatocellular carcinoma patients and controls

    Am. J. Epidemiol.

    (1998)
  • K. Riedel et al.

    Determination of three carcinogenic aromatic amines in urine of smokers and nonsmokers

    J. Anal. Toxicol.

    (2006)
  • N.M. Probst-Hensch et al.

    N-acetyltransferase 2 phenotype but not NAT1∗10 genotype affects aminobiphenyl–hemoglobin adduct levels

    Cancer Epidemiol. Biomarkers Prev.

    (2000)
  • D. Kim et al.

    Cytochrome P450 activation of arylamines and heterocyclic amines

    Annu. Rev. Pharmacol. Toxicol.

    (2005)
  • J. Bendaly et al.

    Role of human CYP1A1 and NAT2 in 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine-induced mutagenicity and DNA adducts

    Xenobiotica

    (2009)
  • P.D. Josephy

    The role of peroxidase-catalyzed activation of aromatic amines in breast cancer

    Mutagenesis

    (1996)
  • W.B. Melchior et al.

    Mutations induced by aromatic amine DNA adducts in pBR322

    Carcinogenesis

    (1994)
  • B.L. Parsons et al.

    Levels of 4-aminobiphenyl-induced somatic H-ras mutation in mouse liver DNA correlate with potential for liver tumor development

    Mol. Carcinog.

    (2005)
  • M.A. Doll et al.

    Codominant expression of N-acetylation and O-acetylation activities catalyzed by N-acetyltransferase 2 in human hepatocytes

    J. Pharmacol. Exp. Ther.

    (2010)
  • S.A. Vaziri et al.

    Variation in enzymes of arylamine procarcinogen biotransformation among bladder cancer patients and control subjects

    Pharmacogenetics

    (2001)
  • Cited by (23)

    • Cigarette smoking and liver diseases

      2022, Journal of Hepatology
      Citation Excerpt :

      Regarding the underlying mechanisms, it is biologically plausible that smoking plays an important role in HCC. Several chemicals present in tobacco smoke, including 4-aminobiphenyl and polycyclic aromatic hydrocarbons, can be metabolised into reactive carcinogenic agents in the liver.96,97 These molecules induce histopathological changes such as liver parenchyma degeneration in animal models.98

    • Primary aromatic amines and cancer: Novel mechanistic insights using 4-aminobiphenyl as a model carcinogen

      2019, Pharmacology and Therapeutics
      Citation Excerpt :

      Based mostly on in vitro studies, the major enzymes that have been thought to catalyze oxidation and conjugation reactions to bioactivate aromatic amines are the cytochrome P450 isoform CYP1A2 followed by one or both of the two arylamine N-acetyltransferases (NATs). Even though studies performed in vivo generally support the role of aromatic amine-DNA adducts in aromatic amine carcinogenesis, they also point to the possibility of alternative bioactivation and tumor promotion/progression pathways for aromatic amines (Kimura et al., 1999; Sugamori et al., 2012). Given this disparity between in vitro and in vivo findings, the goals of this review are: 1) to provide an overview of studies that established the roles of bioactivation enzymes and DNA adducts in aromatic amine carcinogenesis; 2) to highlight studies where levels of aromatic amine DNA adducts have failed to correlate with carcinogenesis; and 3) to present recent studies that may shed light on novel pathways and mechanisms that drive aromatic amine carcinogenesis.

    • Influence of sex and developmental stage on acute hepatotoxic and inflammatory responses to liver procarcinogens in the mouse

      2016, Toxicology
      Citation Excerpt :

      This ‘neonatal’ tumor induction protocol, which involves carcinogen exposure prior to sexual maturity followed by tumor enumeration well into adulthood, provides high sensitivity for genotoxic chemicals (McClain et al., 2001). In previous studies we observed that only male C57BL/6 mice developed liver tumors by one year after postnatal ABP exposure, while levels of the potential ABP-bioactivating enzymes CYP1A2, CYP2E1 and NAT, acute DNA damage (ABP-DNA adducts) and ABP-induced in vivo mutation frequencies were no different between male and female mice (Sugamori et al., 2012; Wang et al., 2015a; Wang et al., 2012). A lack of correlation between ABP bioactivation capacity, DNA adducts and liver tumor incidence has also been observed by others (Chen et al., 2005), and it supports the idea that although DNA damage is necessary for tumor initiation, it may not account for the observed sex difference in ABP-induced tumor growth.

    • Polymorphism p.Val231Ile alters substrate selectivity of drug-metabolizing arylamine N-acetyltransferase 2 (NAT2) isoenzyme of rhesus macaque and human

      2014, Gene
      Citation Excerpt :

      Mice, in particular, have been useful as pharmacological and toxicological models of human NATs, as various “rapid” and “slow” acetylator strains have been characterized (Boukouvala et al., 2002; Erickson et al., 2008; Martell et al., 1991). Moreover, various transgenic mouse models have been generated for NAT (Cornish et al., 2003; Sugamori et al., 2003, 2011), and these have been used to study the carcinogenicity of compounds (Sugamori et al., 2006, 2012). Such engineered strains have also been used to conduct studies on the epigenetic regulation (Wakefield et al., 2010) and developmental expression (Wakefield et al., 2008) of murine Nat genes, advancing our understanding of the endogenous roles of mammalian NATs (Butcher and Minchin, 2012).

    View all citing articles on Scopus
    View full text