Hostname: page-component-7c8c6479df-94d59 Total loading time: 0 Render date: 2024-03-29T11:36:08.529Z Has data issue: false hasContentIssue false

DNA–carcinogen interaction: covalent DNA-adducts of benzo(a)pyrene 7, 8-dihydrodiol 9, 10-epoxides studied by biochemical and biophysical techniques

Published online by Cambridge University Press:  17 March 2009

Astrid Gräslund
Affiliation:
Department of Medical Biophysics, University of Umed, S-901 87 Umeå, Sweden
Bengt Jernström
Affiliation:
Department of Toxicology, Karolinska Institutet, S-104 01 Stockholm, Sweden

Extract

Exposure to various chemicals, either due to occupation or lifestyle, is considered to be a major contributing factor to tumour formation in man (Higginson, 1969; Doll and Peto, 1981). An important and prevalent class of potent carcinogeniccompounds present in the environment is polycyclic aromatic hydrocarbons(PAHs), which are found in various petroleum and combustion products derived from heat and power generation and motor vehicle exhausts (Baum, 1978). Furthermore, since PAHs are generally formed by pyrolysis of organic matters such as tobacco smoking and certain procedures of food preparation, the PAH exposure to humans is extensive.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1989

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adamson, E. D. (1987). Oncogenes in development. Development 99, 449471.Google Scholar
Barbacid, M. (1987). ras genes. A. Rev. Biochem. 56, 779827.CrossRefGoogle ScholarPubMed
Baum, E. J. (1978). Occurrence and surveillance of polycyclic aromatic hydrocarbons. In Polycyclic Hydrocarbons and Cancer, vol. 1 (ed. Gelboin, H. V. and Ts'o, P. O. P.), pp. 4562. New York: Academic Press.Google Scholar
Beland, F. A. (1978). Computer-generated graphic models of the N2-substituted deoxyguanosine adducts of 2-acetylaminofluorene and benzo(a)pyrene and the O6-substituted deoxyguanosine adduct of 1-naphtylamine in the DNA double helix. Chem. biol. Interactions 22, 329339.Google Scholar
Bigger, C. A. H., Sawicki, J. T., Blake, D. M., Raymond, L. G. & Dipple, A. (1983). Products of binding of 7, 2-dimethylbenz(a)anthracene to DNA in mouse skin. Cancer Res. 43, 56475651.Google Scholar
Bishop, J. M. (1983). Cellular oncogenes and retroviruses. A. Rev. Biochem. 53, 301354.Google Scholar
Bohr, V. A., Smith, C. A., Okumoto, D. A. & Hanawalt, P. C. (1985). DNA repair in an active gene: removal of pyrimidine dimers from the DHFR gene of CHO cells is much more efficient than in the genome overall. Cell 40, 359369.Google Scholar
Boles, T. C. & Hogan, M. E. (1984). Site-specific carcinogen binding to DNA. Proc. Natn. Acad. Sci. USA 81, 56235627.Google Scholar
Boles, T. C. & Hogan, M. E. (1986). High-resolution mapping of carcinogen binding sites on DNA. Biochemistry 25, 30393043.Google Scholar
Borgen, A., Darvey, H., Castagnoli, N., Crocker, T. T., Rasmussen, R. E. & Wang, I. Y. (1973). Metabolic conversion of benzo(a)pyrene by Syrian hamster liver microsomes and binding of metabolites to deoxyribonucleic acid. J. med. Chem. 16, 502506.Google Scholar
Bowden, G. T., McGovern, V., Osanna, N. & Rosenthal, H. (1982). Concentration dependent alterations of DNA replication initiation and elongation by benzo(a)pyrene diol epoxide. Carcinogenesis 3, 473480.CrossRefGoogle ScholarPubMed
Bresloff, J. L. & Crothers, D. M. (1975). DNA-ethidium reaction kinetics: demonstration of direct ligand transfer between DNA binding sites. J. molec. Biol. 95, 103123.CrossRefGoogle ScholarPubMed
Brookes, P. & Osborne, M. R. (1982). Mutation in mammalian cells by stereoisomers of anti-benzo(a)pyrene diol epoxide in relation to the extent and nature of the DNA reaction products. Carcinogenesis 3, 12231226.CrossRefGoogle Scholar
Buening, M. K., Wislocki, P. G., Levin, W., Yagi, H., Thakker, D. R., Akagi, H., Koreeda, J., Jerina, D. M. & Conney, A. H. (1978). Tumorigenicity of the optical enantiomers of the diastereomeric benzo(a)pyrene 7, 8-diol 9, 10-epoxides in newborn mice: exceptional activity of (+)-7β, 8α-dihydroxy-9α, 10β-epoxy-7, 8, 9, 10-tetra-hydrobenzo(a)pyrene. Proc. Natn. Acad. Sci. USA 75, 53585361.Google Scholar
Burgess, J. A., Stevens, C. W. & Fahl, W. E. (1985). Mutation at separate gene loci in Salmonella typhimurium TA100 related to DNA nucleotide modification by stereoisomeric benzo(a)pyrene 7, 8-diol 9, 10-epoxides. Cancer Res. 45, 42574262.Google Scholar
Chang, R. L., Wood, A. W., Conney, A. H., Yagi, H., Sayer, J. M., Thakker, D. R., Jerina, D. M. & Levin, W. (1987). Role of diaxial versus diequatorial hydroxyl groups in the tumorigenic activity of a benzo(a)pyrene bay-region diol epoxide. Proc. Natn. Acad. Sci. USA 84, 86338636.CrossRefGoogle ScholarPubMed
Chen, F. M. (1984). Sequence specific binding of tetraols of benzo(a)pyrene-diol-epoxide to DNA in neutral and acidic DNA. Carcinogenesis 5, 753758.Google Scholar
Chen, F. M. (1985). Covalent binding of benzo(a)pyrene diol epoxides to polynucleotides. Biochemistry 24, 50455052.CrossRefGoogle Scholar
Chen, F. M. (1986). Binding of enantiomers of trans-7, 8-dihydroxy-anti-9, 10-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene to polynucleotides. J. biomolec. struct. Dyn. 4, 401418.Google Scholar
Ciejek, E. M., Tsai, M.-J. & O'Malley, R. W. (1983). Actively transcribed genes are associated with the nuclear matrix. Nature 306, 607609.CrossRefGoogle ScholarPubMed
Conney, A. H. (1982). Induction of microsomal enzymes by foreign chemicals and carcinogenesis by polycyclic aromatic hydrocarbons: G. H. A. Clowes Memorial Lecture. Cancer Res. 42, 48754917.Google Scholar
Cooper, C. S., Grover, P. L. & Sims, P. (1983). The metabolism and activation of benzo(a)pyrene. Prog. Drug Metab., 295396.Google Scholar
Cooper, G. M. (1982). Cellular transforming genes. Science 218, 801806.Google Scholar
Cordeiro-Stone, M., Boyer, I.C., Smith, B. A. & Kaufmann, W. K. (1986). Effect of benzo(a)pyrene-diol-epoxide-I on growth of nascent DNA in synchronized human fibroblasts. Carcinogenesis 7, 17751781.Google Scholar
Dandekar, S., Sukumar, S., Zarbl, H., Young, L. J. T. & Cardiff, R. D. (1986). Specific activation of the cellular Harvey-ras oncogene in dimethylbenzanthracene-induced mouse mammary tumours. Mol. Cell Biol. 6, 41044108.Google Scholar
Daudel, P., Duquesne, M., Vigny, P., Grover, P. L. & Sims, P. (1975). Fluorescence spectral evidence that benzo(a)pyrene-DNA products in mouse skin arise from diolepoxides. FEBS Lett. 57, 250253.Google Scholar
Day, R. S. III, Scudiero, D. & Dimattina, M. (1978). Excision repair by human fibroblasts of DNA damaged by r-7, t-8-dihydroxy-t-9, 10-oxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene. Mut. Res. 50, 383394.Google Scholar
DiGiovanni, J., Sawyer, T. W. & Fisher, E. P. (1986). Correlation between formation of a specific hydrocarbon-deoxyribonucleoside adduct and tumor initiating activity of 7, 12-dimethylbenz(a)anthracene and its 9- and 10-monofluoro derivatives in mice. Cancer Res. 46, 43364341.Google Scholar
Dipple, A. (1985). Polycyclic aromatic hydrocarbon carcinogenesis: an introduction. In Polycyclic Hydrocarbons and Carcinogenesis, ACS Symposium Series 283 (ed. Harvey, R. G.), pp. 117. Washington, D.C.: American Chemical Society.Google Scholar
Dipple, A., Sawicki, J. T., Moschel, R. C. & Bigger, C. A. H. (1983). 7, 12-dimethylbenz(a)anthracene–DNA interactions in mouse embryo cell cultures and mouse skin. In Extrahepatic Drug Metabolism and Chemical Carcinogenesis (ed. Rydström, J., Montelius, J. and Bengtsson, M.), pp. 439448. Amsterdam: Elsevier.Google Scholar
Dock, L., Undeman, O., Gräslund, A. & Jernström, B. (1978). Fluorescence study of DNA complexes formed after metabolic activation of benzo(a)pyrene derivatives. Biochem. biophys. Res. Commun. 85, 12751282.Google Scholar
Doll, R. & Peto, R. (1981). The causes of cancer: Quantitative estimates of avoidable risks of cancer in the United States today. J. natn. Cancer Inst. 66, 11911308.CrossRefGoogle ScholarPubMed
Eriksson, M., Eriksson, S., Jernström, B., Nordén, B. & Gräslund, A. (1988 c). Excimer fluorescence from (+)-anti-benzo(a)pyrene diol epoxide covalently bound to poly(dG-dC). Structural implications (to be published).Google Scholar
Eriksson, M., Jernström, B., Gräslund, A. & Nordén, B. (1986). Binding geometry of benzo(a)pyrene diol epoxide covalently bound to DNA. A flow linear dichroism study. J. Chem. Soc., Chem. Commun. 16131615.CrossRefGoogle Scholar
Eriksson, M., Nordén, B., Jernström, B. & Gräslund, A. (1988 a). Binding geometries of benzo(a)pyrene diol epoxide isomers covalently bound to DNA. Orientational distribution. Biochemistry 27, 12131221.CrossRefGoogle ScholarPubMed
Eriksson, M., Nordén, B., Jernström, B. & Gräslund, A. (1988 d). NaCl induced B to Z transition in poly(dG-dC) modified with benzo(a)pyrene diol epoxides. Polarized light spectroscopy studies (to be published).Google Scholar
Eriksson, M., Nordén, B., Jernström, B., Gräslund, A. & Lycksell, P.-O. (1988 b). Observation of excimer formation in the covalent adducts of 9, 10-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene-7, 8-diol with poly(dG-dC). J. Chem. Soc., Chem. Commun. 211212.CrossRefGoogle Scholar
Evaluation of Carcinogenic Risk (1972). International Agency for Research on Cancer Monographs, vol. 3.Google Scholar
Evaluation of Carcinogenic Risk (1983). International Agency for Research on Cancer Monographs, vol. 32.Google Scholar
Geacintov, N. E. (1985). Mechanisms of interaction of polycyclic aromatic diol epoxides with DNA and structures of the adducts. In Polycyclic Hydrocarbons and Carcinogenesis, ACS Symptosium Series 293 (ed. Harvey, R. G.), pp. 107142. Washington, D.C.: American Chemical Society.Google Scholar
Geacintov, N. E. (1986 a). Mechanisms of reaction of polycyclic aromatic epoxide derivatives with nucleic acids. In Polycyclic Aromatic Hydrocarbon Carcinogenesis: Structures Activity Relationships (ed. Yang, S. and Silverman, B. D.), pp. 181206. Boca Raton: CRC Press.Google Scholar
Geacintov, N. E. (1986 b). Is intercalation a critical factor in the covalent binding of mutagenic and tumorigenic polycyclic aromatic diol epoxides to DNA? Carcinogenesis 7, 759766.Google Scholar
Geacintov, N. E., Carberry, S. E., Swenberg, C. E. & Harvey, R. G. (1988). Formation of adducts of polycyclic aromatic diol epoxide derivatives with supercoiled DNA investigated by kinetic flow linear dichroism techniques. Poster presented at the 79th Annual Meeting of the American Association for Cancer Research, New Orleans, May, 1988.Google Scholar
Geacintov, N. E., Gagliano, A., Ivanovic, V. & Weinstein, I. B. (1978). Electric linear dichroism study of the orientation of benzo(a)pyrene-7, 8-dihydrodiol 9, 10-oxide covalently bound to DNA. Biochemistry 17, 52565262.Google Scholar
Geacintov, N. E., Gagliano, A. G., Ibanez, V. & Harvey, R. G. (1982). Spectroscopic characterizations and comparisons of the structures of the covalent adducts derived from the reactions of 7, 8-dihydroxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene-9, 10-oxide, and the 9, 10-epoxides of 7, 8, 9, 10-tetrahydrobenzo(a)pyrene and 9, 10, 11, 12-tetrahydrobenzo(e)pyrene with DNA. Carcinogenesis 3, 247253.Google Scholar
Geacintov, N. E., Ibanez, V., Gagliano, A. G., Jacobs, S. A. & Harvey, R. G. (1984). Stereoselective covalent binding of anti-benzo(a)pyrene diol epoxide to DNA. Conformation of enantiomer adducts. J. Biomolec. Struct. Dyn. 1, 14731484.Google Scholar
Geacintov, N. E., Ibanez, V., Gagliano, A., Yoshida, H. & Harvey, R. G. (1980). Kinetics of hydrolysis to tetraols and binding of benzo(a)pyrene-7, 8-dihydrodiol-9, 10-oxide and its tetraol derivatives of DNA. Conformation of adducts. Biochem. biophys. Res. Commun. 92, 13351342.Google Scholar
Geacintov, N. E., Yoshida, H., Ibanez, V. & Harvey, R. G. (1981). Non-covalent intercalative binding of 7, 8-dihydroxy-9, 10-epoxybenzo(a)pyrene to DNA. Biochem. biophys. Res. Commun. 100, 15691577.Google Scholar
Geacintov, N. E., Zinger, D., Ibanez, V., Santella, R., Grunberger, D. & Harvey, R. G. (1987). Properties of covalent benzo(a)pyrene diol epoxide-DNA adducts investigated by fluorescence techniques. Carcinogenesis 8, 925935.Google Scholar
Gelboin, H. V. (1980). Benzo(a)pyrene metabolism, activation and carcinogenesis: role and regulation of mixed functions. Physiol. Rev. 60, 11071166.CrossRefGoogle Scholar
Glatt, H., Seidel, A., Bochnitschek, W., Marquardt, H., Marquardt, H., Hodgson, R. M., Grover, P. L. & Oesch, F. (1986). Mutagenic and cell transforming activities of triol-epoxides as compared to other chrysene metabolites. Cancer Res. 46, 45564565.Google Scholar
Guéron, M., Kochoyan, M. & Leroy, J. L. (1987). A single mode of DNA base-pair opening drives imino proton exchange. Nature 328, 8992.Google Scholar
Hancock, R. (1982). Topological organization of interphase DNA: the nuclear matrix and other skeletal structures. Biol. Cell 46, 102122.Google Scholar
Hancock, R. & Hughes, M. E. (1982). Organisation of DNA in the interphase nucleus. Biol. Cell 44, 201212.Google Scholar
Harris, C. C. (1987). Human tissues and cells in carcinogenesis research. Cancer Res. 47. 110.Google Scholar
Harvey, R. G. & Geacintov, N. E. (1988). Intercalation and binding of carcinogenic hydrocarbon metabolites to nucleic acids. Acc. Chem. Res. 21, 6673.Google Scholar
Haseltine, W. A., Lo, K. M. & D'Andrea, A. D. (1980). Preferred sites of strand scissions in DNA modified by anti-diol epoxide of benzo(a)pyrene. Science 209, 929931.Google Scholar
Higginson, I. (1969). Present trends in cancer epidemiology. Can. Cancer Conf. 8, 4075.Google ScholarPubMed
Hingerty, B. E. & Broyde, S. (1985). Carcinogen–base stacking and base–base stacking in dCpdG modified by (+) and (−)-anti-BPDE. Biopolymers 24, 22792299.CrossRefGoogle ScholarPubMed
Hogan, M. E., Dattagupta, N. & Whitlock, J. P. Jr., (1981). Carcinogen-induced alteration of DNA structure. J biol. Chem. 256, 45044513.CrossRefGoogle ScholarPubMed
Ibanez, V., Geacintov, N. E., Gagliano, A. G., Brandimarte, S. & Harvey, R. G. (1980). Physical binding of tetraols derived from 7, 8-dihydroxy-9, 10-epoxy benzo(a)pyrene to DNA. J. Am. chem. Soc. 102, 56615666.Google Scholar
Igo-Kemenes, T., Hörz, W. & Zachau, H. G. (1982). Chromatin. A. Rev. Biochem. 51, 89121.Google Scholar
Ivanovic, V., Geacintov, N. E. & Weinstein, I. B. (1976). Cellular binding of benzo(a)pyrene to DNA characterized by low temperature fluorescence. Biochem. biophys. Res. Commun. 70, 11721179.CrossRefGoogle ScholarPubMed
Jack, P. L. & Brookes, P. (1982). Mechanism for the loss of preferential benzo(a)pyrene binding to the linker DNA of chromatin. Carcinogenesis 3, 341344.Google Scholar
Jerina, D. M. & Lehr, R. E. (1978). The bay-region theory: a quantum mechanical approach to aromatic hydrocarbon-induced carcinogenicity. In Microsomes and Drug Oxidations (ed. Ulrich, V., Roots, I., Hilderbrandt, A. and Estabrook, R. W.), pp. 709720. Elmsford, New York: Pergamon Press.Google Scholar
Jerina, D. M., Sayer, J. M., Thakker, D. R. & Yagi, H. (1980). Carcinogenicity of polycyclic aromatic hydrocarbons: the bay region theory. In Carcinogenesis: Fundamental Mechanisms and Environmental Effects (ed. Pullman, B., Ts'o, P. O. P. and Gelboin, H.), pp. 112. Reidel Publishing Co.Google Scholar
Jernström, B., Lycksell, P.-O., Gräslund, A., Ehrenberg, A. & Nordén, B. (1983). Spectroscopic properties of benzo(a)pyrene-7, 8-dihydrodiol-9, 10-epoxides covalently bound to DNA. In Extrahapatic Drug Metabolism and Chemical Carcinogenesis (ed. Rydström, J., Montelius, J. and Bengtsson, M.), pp. 469477. Amsterdam: Elsevier.Google Scholar
Jernström, B., Lycksell, P.-O., Gräslubd, A. & Nordén, B. (1984). Spectroscopic studies of DNA complexes formed after reaction with anti-benzo(a)pyrene-7, 8-dihydrodiol-9, 10-oxide enantiomers of different carcinogenic potency. Carcinogenesis 5, 11291135.CrossRefGoogle Scholar
Jernström, B., Orrenius, S., Undeman, O., Gräslund, A. & Ehrenberg, A. (1978). Fluorescence study of DNA-binding metabolites of benzo(a)pyrene in hepatocytes isolated from 3-methylcholanthrene-treated rats. Cancer Res. 38, 26002607.Google Scholar
Jovin, T. M., McIntosh, L. P., Arndt-Jovin, D. M., Zarling, D. A., Robert-Nicoud, M., van de Sande, J. M., Jörgenson, K. F. & Eckstein, F. (1983). Left-handed DNA: from synthetic polymers to chromosomes. J. biomolec. Struct. Dyn. 1, 2156.Google Scholar
Kaneko, M. V. & Cerutti, P. A. (1982). Excision of benzo(a)pyrene diol epoxide I adducts from nucleosomal DNA of confluent normal human fibroblasts. Chem. biol. Interactions 38, 261274.Google Scholar
Kaufmann, W. K., Boyer, J. C., Smith, B. A. & Cordeiro-Stone, M. (1985). DNA repair and replication in human fibroblasts treated with (±)-r-7, t-8-dihydroxy-t-9, 10-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene. Biochim. biophys. Acta 824, 146151.CrossRefGoogle Scholar
Kim, S. K., Geacintov, N. E., Zinger, D. & Sutherland, J. C. (1988). Fluorescence spectral characteristics and fluorescence decay profiles of covalent polycyclic aromatic carcinogen–DNA adducts. In Brookhaven Symposium in Biology, No. 35. Synchrotron Radiation in Structural Biology. Upton, N.Y.Google Scholar
Klein, G. & Klein, E. (1984). Oncogene activation and tumor progression. Carcinogenesis 5, 429435.Google Scholar
Kolubayev, V., Brenner, H. C. & Geacintov, N. E. (1987). Stereoselective covalent binding of enantiomers of anti-benzo(a)pyrene diol epoxide to DNA as probed by optical detection of magnetic resonance. Biochemistry 26, 26382641.CrossRefGoogle ScholarPubMed
Kootstra, A. (1986). The dynamics of chromatin carcinogen interactions in the human cell. Nucl. Acid Res. 14, 98979909.Google Scholar
Koostra, A. & Slaga, T. J. (1980). Binding of isomers of benzo(a)pyrene diol-epoxide to chromatin. Biochem. biophys. Res. Commun. 93, 954959.Google Scholar
Koostra, A., Slaga, T. J. & Olins, D. E. (1979). Interaction of benzo(a)pyrene diolepoxide with nuclei and isolated chromatin. Chem. biol. Interactions 28, 225236.CrossRefGoogle Scholar
LeBreton, P. R. (1985). The intercalation of benzo(a)pyrene and 7, 12-dimethylbenz(a)anthracene metabolites and metabolite model compounds in DNA. In Polycyclic Hydrocarbons and Carcinogenesis. ACS Symposium Series 283 (ed. Harvey, R. G.), pp. 209238. Washington, D.C.: American Chemical Society.Google Scholar
Lehr, R. E., Kumar, S., Levin, W., Wood, A. W., Chang, R. L., Conney, A. H., Yagi, H., Sayer, J. M. & Jerina, D. M. (1985). The bay-region theory of polycyclic aromatic hydrocarbon carcinogenesis. In Polycyclic Hydrocarbons and Carcinogenesis, ACS Symposium Series 283 (ed. Harvey, R. G.), pp. 6484. Washington, D.C.: American Chemical Society.Google Scholar
Lobanenkov, V. V., Plumb, M., Goodwin, G. H. & Grover, P. L. (1986). The effect of neighbouring bases on G-specific DNA cleavage mediated by treatment with the anti-diol epoxide of benzo(a)pyrene in vitro. Carcinogenesis 7, 16891695.Google Scholar
Lycksell, P.-O., Gräslund, A., Claesens, F., McLaughlin, L. W., Larsson, U. & Rigler, R. (1987). Base pair opening dynamics of a 2-aminopurine substituted Eco RI restriction sequence and its unsubstituted counterpart in oligonucleotides. Nucl. Acids Res. 15, 90119025.Google Scholar
Lycksell, P.-O., Gräslund, A., Ehrenberg, A., Jernström, B. & Nordén, B. (1985). Facilitation of the conversion from B to Z conformation in poly(dG–dC) modified by anti-benzo(a)pyrene-7, 8-dihydrodiol-9, 10-epoxide. J. biol. Chem. 260, 58575859.CrossRefGoogle Scholar
MacLeod, M. C. & Selkirk, J. K. (1982). Physical interactions of isomeric benzo(a)pyrene diol-epoxides to DNA. Carcinogenesis 3, 287292.Google Scholar
Marshall, C. J., Vousden, K. H. & Phillips, D. H. (1984). Activation of c-Ha-ras-i proto-oncogene by in vitro modification with a chemical carcinogen, benzo(a)pyrene diol-epoxide. Nature 310, 586589.CrossRefGoogle Scholar
Meehan, T., Gamper, H. & Becker, J. F. (1982). Characterization of reversible physical binding of benzo(a)pyrene derivatives to DNA. J. Biol. Chem. 257, 1047910485.CrossRefGoogle ScholarPubMed
Meehan, T. & Straub, K. (1979). Double-stranded DNA stereoselectivity binds benzo(a)pyrene diol epoxides. Nature 277, 410412.Google Scholar
Miller, E. C. (1978). Some current perspectives on chemical carcinogenesis in humans and experimental animals: presidential address. Cancer Res. 38, 14791496.Google ScholarPubMed
Miller, J. A. (1970). Carcinogenesis by chemicals: an overview – G. H. A. Clowes memorial lecture. Cancer Res. 30, 559576.Google Scholar
Miller, K. J., Taylor, E. R. & Dommen, J. (1985). A mechanism for the stereoselectivity and binding of benzo(a)pyrene diol epoxides to DNA. In Polycyclic Hydrocarbons and Carcinogenesis, ACS Symposium Series 283 (ed. Harvey, R. G.), pp. 239288. Washington, D.C.: American Chemical Society.Google Scholar
Mirau, P. A. & Kearns, D. R. (1983). The effect of intercalating drugs on the kinetics of the B to Z transition of poly(dG-dC). Nucl. Acids Res. 11, 19311941.CrossRefGoogle Scholar
Mironov, N. M., Grover, P. L. & Sims, P. (1983). Preferential binding of polycyclic hydrocarbons to matrix-bound DNA in rat-liver nuclei. Carcinogenesis 4, 189193.Google Scholar
Moussaoui, K., Geacintov, N. E. & Harvey, R. G. (1985). Reactivity and binding of benzo(a)pyrene diol epoxide to poly(dG–dC)·(dG–dC) and poly(dG–m5dC)(dG–m5dC) in the B and Z forms. Biophys. Chem. 22, 285297.Google Scholar
Nicholson, G. L. (1987). Tumor cell instability, diversification, and progression to the metastatic phenotype: from oncogene to oncofetal expression. Cancer Res. 47, 14731487.Google Scholar
Nishimura, S. & Sekiya, T. (1987). Human cancer and cellular oncogenes. Biochem. J. 243, 313327.Google Scholar
Nordén, B. (1978). Applications of linear dichroism spectroscopy. Appl. Spectrosc. Rev. 14, 157248.CrossRefGoogle Scholar
Nordheim, A. & Rich, A. (1983). Negatively supercoiled simian virus-40 DNA contains Z-DNA segments within transcriptional enhancer sequences. Nature 303, 674679.Google Scholar
Obi, F. O., Ryan, A. J. & Billet, M. A. (1986). Preferential binding of the carcinogen benzo(a)pyrene to DNA in active chromatin and the nuclear matrix. Carcinogenesis 7, 907913.Google Scholar
Osborne, M. R., Jacobs, S., Harvey, R. G. & Brookes, P. (1981). Minor products from the reaction of (+) and (−) benzo(a)pyrene-anti-diolepoxide with DNA. Carcinogenesis 2, 553558.Google Scholar
Osborne, M. & Merrifield, K. (1985). Depurination of benzo(a)pyrene-diolepoxide treated DNA. Chem. biol. Interactions 53, 183195.Google Scholar
Pardoll, D. M., Vogelstein, B. & Coffey, D. S. (1980). A fixed site of DNA replication in eukaryotic cells. Cell 19, 527536.Google Scholar
Pearl, L. H. & Neidle, S. (1986). Origins of stereospecificity in DNA by antibenzo(a)pyrene diol epoxides. A molecular modelling study. FEBS Lett. 209, 269276.Google Scholar
Pelling, J. C., Ernst, S. M., Strawhecker, J., Johnson, J., Nairn, R. & Slaga, T. J. (1986). Elevated expression of Ha-ras is an early event in two-stage skin carcinogenesis in SENCAR mice. Carcinogenesis 7, 15991602.Google Scholar
Pelling, J. C., Fischer, S. M., Neades, R., Strawhecker, J. & Schweickert, L. (1987). Elevated expression and point mutation of the Ha-ras proto-oncogene in mouse skin tumours promoted by benzoyl peroxide and other promoting agents. Carcinogenesis 8, 14811484.CrossRefGoogle ScholarPubMed
Pelling, J. C., Neades, R. & Strawhecker, J. (1988). Epidermal papillomas and carcinonomas induced in uninitiated mouse skin by tumor promoters alone contain a point mutation in the 6ist codon of the Ha-ras oncogene. Carcinogenesis 9, 665667.Google Scholar
Pelling, J. C. & Slaga, T. J. (1982). Comparison of levels of benzo(a)pyrene diolepoxide diastereomers covalently bound in vitro to macromolecular components of the whole epidermis versus the basal cell layer. Carcinogenesis 3, 11351141.CrossRefGoogle Scholar
Pelling, J. C., Slaga, T. J. & DiGiovanni, J. (1984). Formation and persistence of DNA, RNA, and protein adducts in mouse skin exposed to pure optical enantiomers of 7β, 8α dihydroxy-9α, 10α-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene in vivo. Cancer Res. 44, 10811086.Google Scholar
Phillips, D. H. (1983). Fifty years of benzo(a)pyrene. Nature 303, 468472.Google Scholar
Pohl, F. M. (1987). Left-handed DNA: energetic and dynamic aspects. In Structure, Dynamics and Function of Biomolecules (ed. Ehrenberg, A., Rigler, R., Gräslund, A. and Nilsson, L.), pp. 224228. Heidelberg: Springer Verlag.Google Scholar
Pohl, F. M. & Jovin, T. M. (1972). Salt-induced co-operative conformational change of a synthetic DNA: equilibrium and kinetic studies with poly(dG–dC). J. molec. Biol. 67. 375396.Google Scholar
Prusik, T., Geacintov, N. E., Tobiasz, C., Ivanovic, V. & Weinstein, I. B. (1979). Fluorescence study of the physico-chemical properties of a benzo(a)pyrene 7, 8 dihydrodiol 9, 10-oxide derivative bound covalently to DNA. Photochem. Photobiol. 29, 223232.Google Scholar
Pulkrabek, P., Leffler, S., Grunberger, D. & Weinstein, I. B. (1979). Modification of deoxyribonucleic acid by a diol epoxide of benzo(a)pyrene. Relation to deoxyribonucleic acid structure and conformation and effects on transfectional activity. Biochemistry 16, 31273132.Google Scholar
Pulkrabek, P., Leffler, S., Weinstein, I. B. & Grunberger, D. (1977). Conformation of DNA modified with a dihydrodiol epoxide derivative of benzo(a)pyrene. Biochemistry (Wash.) 16, 31273132.Google Scholar
Pullman, A. & Pullman, B. (1955). Electronic structure and carcinogenic activity of aromatic molecules. In Advances in Cancer Research, vol. III (ed. Greenstein, J. P. and Haddow, A.), pp. 117169. New York: Academic Press.Google Scholar
Quintanilla, M., Brown, K., Ramdsen, M. & Balmain, A. (1986). Carcinogenspecific mutation and amplification of Ha-ras during mouse skin carcinogenesis. Nature 322, 7880, 1986.Google Scholar
Ramstein, J., Ehrenberg, M. & Rigler, R. (1980). Fluorescence relaxation of proflavin-deoxyribonucleic acid interaction. Kinetic properties of a base-specific reaction. Biochemistry 19, 39383948.Google Scholar
Reddy, E. P., Reynolds, R. K., Santos, E. & Barbacid, M. (1982). A point mutation is responsible for the acquisition of transforming properties by the T24 human bladder carcinoma oncogene. Nature 300, 149152.Google Scholar
Regan, J. D., Francis, A. A., Dunn, W. C., Hernandez, O., Yagi, H. & Jerina, D. M. (1978). Repair of DNA damaged by mutagenic metabolites of benzo(a)pyrene in human cells. Chem. biol. Interactions 20, 279287.Google Scholar
Rich, A., Nordheim, A. & Wang, A. M.-J. (1984). The chemistry and biology of lefthanded Z-DNA. A. Rev. Biochem. 53, 791846.CrossRefGoogle ScholarPubMed
Roberts, J. D. & Caserio, M. C. (1965). Basic Principles of Organic Chemistry. New York, Amsterdam: W. A. Benjamin, Inc.Google Scholar
Sage, E. & Haseltine, W. A. (1984). High ratio of alkali-sensitive lesions to total DNA modification induced by benzo(a)pyrene diol epoxide. J. biol. Chem. 259, 1109811102.Google Scholar
Seidman, M., Slor, H. & Bustin, M. (1983). The binding of a carcinogen to the nucleosomal and non-nucleosomal regions of the simian virus 40 chromosome in vivo. J. biol. Chem. 258, 52155220.CrossRefGoogle Scholar
Shih, C., Shilo, B.-Z., Goldfarb, M. P., Dannenberg, A. & Weinberg, R. A. (1979). Passage of phenotypes of chemically transformed cells via transfection of DNA and chromatin. Proc. Natn. Acad. Sci. USA 76, 57145718.Google Scholar
Shinohara, K. & Cerutti, P. A. (1977). Excision repair of benzo(a)pyrenedeoxyguanosine adducts in baby hamster kidney 21/C13 cells and in secondary mouse embryo fibroblasts C57/BL/6J. Proc. Natn. Acad. Sci. USA 74, 979983.CrossRefGoogle Scholar
Sims, P., & Grover, P. L., (1974). Epoxides in polycyclic aromatic hydrocarbon metabolism and carcinogenesis. Adv. Cancer Res. 20, 165275.CrossRefGoogle ScholarPubMed
Sims, P., Grover, P. L., Swaisland, A., Pal, K. & Hewer, A. (1974). Metabolic activation of benzo(a)pyrene proceeds by a diol epoxide. Nature 252, 326328.Google Scholar
Slaga, T. J. (ed.) (1984) Mechanisms of tumor promotion, vol. II. Boca Raton: CRC Press.Google Scholar
Slaga, T. J., Bracken, W. J., Gleason, G., Levin, W., Yagi, H., Jerina, D. M. & Conney, A. H. (1979). Marked differences in the skin tumor-initiating activities of the optical enantiomers of the diastereomeric benzo(a)pyrene 7, 8-diol 9, 10-epoxides. Cancer Res. 39, 6771.Google Scholar
Stevens, C. W., Bouck, N., Burgess, J. A. & Fahl, W. E. (1985). Benzo(a)pyrene diolepoxides: different mutagenic efficiency in human and bacterial cells. Mut. Res. 152, 514.Google Scholar
Stevens, C. W., Manoharan, T. H. & Fahl, W. E. (1988). Characterization of mutagen-activated cellular oncogenes that confer anchorage independence to human fibroblasts and tumorigenicity to NIH 3T3 cells: sequence analysis of an enzymatically amplified mutant HRAS allele. Proc. Natn. Acad. Sci. USA 85, 38753879.CrossRefGoogle ScholarPubMed
Stezowski, J. J., Joos-Guba, G., Schönwälder, K. H., Straub, A. & Glusker, J. P. (1987). Preparation and characterization in solution of polynucleotides alkylated by activated carcinogenic polycyclic aromatic hydrocarbons. Poster presented at the 9th International Biophysics Congress, Jerusalem, 1987.Google Scholar
Tabin, C. J., Bradley, S. M., Bargmann, C. I., Weinberg, R. A., Papageorge, A. G., Scolnick, E. M., Dhar, R., Lowy, D. R. & Chang, E. H. (1982). Mechanism of activation of a human oncogene. Nature 300, 143149.Google Scholar
Thakker, D. R., Yagi, H., Levin, W., Wood, A. W., Conney, A. H. & Jerina, D. M. (1985). Polycyclic aromatic hydrocarbons: metabolic activation to ultimate carcinogens. In Bioactivation of Foreign Compounds (ed. Anders, M. W.), pp. 177242. Academic Press.Google Scholar
Tsang, W.-S. & Griffin, G. W. (1981). Metabolic Activation of Poly nuclear Aromatic Hydrocarbons. Pergamon Press.Google Scholar
Undeman, O., Lycksell, P.-O., Gräslund, A., Astlind, T., Ehrenberg, A., Jernstrom, B., Tjerneld, F. & Nordén, B. (1983). Covalent complexes of DNA and two stereoisomers of benzo(a)pyrene-7, 8-dihydrodiol-9, 10-epoxide studied by fluorescence and linear dichroism. Cancer Res. 43, 18511860.Google Scholar
Undeman, O., Sahlin, M., Gräslund, A., Ehrenberg, A., Dock, L. & Jernstrom, B. (1980). Fluorescence study of (±)-trans-7, 8-dihydroy-7, 8-dihydro-benzo(a)pyrene metabolism in vitro and binding to DNA. Biochem. Biophys. Res. Comm. 94, 458546.CrossRefGoogle Scholar
Vousden, K. H., Bos, J. L., Marshall, C. J. & Phillips, D. H. (1986). Mutations activating human c-Ha-ras 1 proto-oncogene (HRAS1) induced by chemical carcinogens and epurination. Proc. Natn. Acad. Sci. USA 83, 12221226.CrossRefGoogle Scholar
Weinberg, R. A. (1985). The action of oncogenes in the cytoplasm and nucleus. Science 230, 770776.Google Scholar
Weinstein, I. B., Jeffrey, A. M., Leffler, S., Pulkrabek, P., Yamasaki, H. & Grunberger, D. (1978). Interactions between polycyclic aromatic hydrocarbons and cellular macromolecules. In Polycyclic Hydrocarbons and Cancer, vol. II (eds. Gelboin, H. V. and Ts'o, P. O. P.), pp. 436, New York: Academic Press.Google Scholar
Wolfe, A., Shimer, G. H. Jr. & Meehan, T. (1987). Polycyclic aromatic hydrocarbons physically intercalate into duplex regions of denatured DNA. Biochemistry 26, 63926396.CrossRefGoogle ScholarPubMed
Wood, A. W., Chang, L. R., Levin, W., Yagi, H., Thakker, D. R., Jerina, D. M. & Conney, A. H. (1977). Differences in mutagenicity of the optical enantiomers of the diastereomeric benzo(a)pyrene 7, 8-diol-9, 10-epoxides. Biochem. biophys. Res. Commun. 77, 13891396.Google Scholar
Yamanishi, D. T., Bowden, G. T. & Cress, A. E. (1987). An analysis of DNA replication in synchronized CHO cells treated with benzo(a)pyrene diol epoxide. Biochim. biophys. Acta 910, 3442.Google Scholar
Yamasaki, H., Roush, T. W. & Weinstein, I. B. (1978). Benzo(a)pyrene 7, 8-dihydrodiol 9, 10-oxide modification of DNA: relation to chromatin structure and reconstitution. Chem. biol. Interactions 23, 201213.CrossRefGoogle ScholarPubMed
Yang, L. L., Maher, V. M. & McCormick, J. J. (1980). Error-free excision of the cytotoxic, mutagenic N2-deoxyguanosine DNA adduct formed in human fibroblasts by (±)-7β, 8α-dihydroxy-9α, 10-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene. Proc. Natn. Acad. Sci. USA 77, 59335937.Google Scholar
Yang, S. K., Mushtaq, M. & Chiu, P.-L. (1985). Stereoselective metabolism and activations of polycyclic aromatic hydrocarbons. In Polycyclic Hydrocarbons and Carcinogenesis, ACS Symposium Series 283 (ed. Harvey, R. G.), pp. 1034. Washington, D.C.: American Chemical Society.Google Scholar
Zegar, I., Lycksell, P.-O., Gräslund, A., Eriksson, M., Nordén, B. & Jernström, B. (1987). Spectroscopic studies on double stranded poly[d(G–C)·(G–C)] in B and Z forms after covalent modification with the anti isomer of trans-7, 8-dihydroxy-9, 10-epoxy-7, 8, 9, 10-tetrahydrobenzo(a)pyrene. Carcinogenesis 8, 899905.Google Scholar
Zinger, D. (1986). Fluorescence studies of the conformations and photostabilities of the covalent adducts of benzo(a)pyrene diol epoxide and DNA. Ph.D. Dissertation, New York University, New York.Google Scholar
Zinger, D., Geacintov, N. E. & Harvey, R. G. (1987). Conformations and selective photodissociation of heterogeneous benzo(a)pyrene diol epoxide enantiomer–DNA adducts. Biophys. Chem. 27, 131138.Google Scholar