23‘?qu09 STATE UN VERSITY LIBPAWES ”‘T‘fifl l ”M. l: ‘ll‘ ll ll ‘5}‘1‘ W l‘ ' mm lulllmxlggllllllll . Mkhigan State University This is to certify that the dissertation entitled MUTATIONAL ANALYSIS OF THE H-ras ONCOCENE IN SPONTANEOUS B6C F MOUSE LIVER TUMORS AND TUMORS INDUCED WITH A aENOTOXIC AND NONGENOTOXIC HEPATOCARCINOGENS presented by Tony R. Fox has been accepted towards fulfillment of the requirements for . M Ph D degree in icrobiology sci/M ‘ as w Major professor Date MSU is an Affirmative Action/Equal Opportunity Institution 042771 .4 _ ‘ --, f4 7, J- PLACE IN RETURN BOX to remove thle checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Adieu/Equal Opportunity lrltltution MUTATIONAL ANALYSIS OF THE H-ms ONCOCENE IN SPONTANEOUS B5C3F1 MOUSE LIVER TUMORS AND TUMORS INDUCED WITH A GENOTOXIC AND NONGENOTOXIC HEPATOCARCINOGENS By Tony R. Fox A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Microbiology c:_‘ LOOSGIEI ABSTRACT MUTATIONAL ANALYSIS OF THE H-ras ONCOCENE IN SPONTANEOUS BsC3F1 MOUSE LIVER TUMORS AND TUMORS INDUCED WITH A GENOTOXIC AND N ONGENOTOXIC HEPATOCARCINOGENS BY Tony R. Fox The frequency and mutational profile of H-ras gene activation was determined in spontaneous liver tumors of male BéC3F1 mice and in tumors induced with the genotoxic hepatocarcinogen benzidine-ZHCI or the nongenotoxic hepatocarcinogens phenobarbital, chloroform or ciprofibrate. DNA sequence analysis of the H-ras gene from representative tumors revealed that 32 of 50 (64%) spontaneous tumors and 13 of 22 (59%) benzidine-ZHCl induced tumors contained a point mutation in codon 61. Tumors induced with the nongenotoxic agents had a much lower frequency of codon 61 mutations, i.e., phenobarbital 1 of 15 (7%), chloroform 5 of 24 (21%) and ciprofibrate 8 of 39 (21 %). No mutations were observed at codons 12, 13 or 117 in tumors from any of the groups. Only three base pair substitutions within codon 61 were found. The one most frequently detected in all of the groups was a CC to AT transversion at the lst nucleotide position, occurring at a 59%, 85%, 100%, 80%, and 88%, frequency in the spontaneous tumors and in the tumors induced with benzidine 2-HC1, phenobarbital, chloroform and ciprofibrate respectively. In these same groups a A-T to CC transition or a A-T to T-A transversion at the second nucleotide position occurred at a frequency of 34%, 8%, 0%, 0%, 12% and 6%, 8%, 0%, 20%, 0% respectively. The number of tumors carrying an activated H-ras gene in the nongenotoxic treatment groups is within the range that would be expected if those animals had not received any treatment. This indicates that the activation of the H-ras gene in those tumors is probably the result of a spontaneous event. The data suggest that these toxicologically and pharmacologically diverse nongenotoxic hepatocarcinogens increase the frequency of liver tumors but do not induce mutations in the H—ras gene. Instead these agents appear to interact with a population of cells that do not contain an activated H-ras gene. This suggests that the mechanisms of tumor development by these nongenotoxic carcinogens differs at least partially from the mechanisms responsible for the development of spontaneous tumors or those induced by a typical genotoxic agent. Dedicated to the memory of my father, Robert L. Fox, who recognized and kindled my natural curiosity of the unknown and showed me how knowledge could provide the means to satisfy it; and to my mother, Doris L. Fox, for her love, guidance and support over the years. Acknowledgements I wish to express my appreciation to Dr. I. Justin McCormick for serving as the director of this research and also thank Dr. Veronica M. Maher for her encouragement and support. I would also like to thank the other members of my graduate committee: Dr. Alan Schumann, Dr. Susan Conrad and Dr. Leeland Velicer for their invaluable time. A special debt of gratitude is extended to Dr. Iia-ling Yang and Dr. Peggy Schott for their technical advice on DNA sequencing and DNA oligonucleotide synthesis respectively. I would also like to thank Sherry Pagels for her excellent technical assistance. I am especially grateful to Dr. Philip Watanabe for providing me with a leave of absence from my job so that I could retum to the university in order to fulfill the requirements for this degree and to The Dow Chemical Company for their financial support. I would also like to thank Mrs. Gywen Knight for her gracious hospitality while I lived in her home during my stay in East Lansing. I owe an unrepayable debt of appreciation to my wife, Terra D. Fox, for here love, patience, understanding and enduring support. This was especially important during the year and a half that we lived apart while I stayed in East Lansing and she maintained our home in Midland. iii TABLE OF CONTENTS LIST OF TABLES vi LIST OF FIGURES vii INTRODUCTION 1 CHAPTER I. LITERATURE REVIEW 7 A. The Process of Carcinogenesis 7 1, Initiation 7 2. Promotion 9 3. Progression 12 4. Properties of benign and malignant mmm-c 1'; 5. Tumor metastasis 18 6. Tumor heterogeneity 19 B, Chemical Cardnngpnmie 77 1. Genotoxic carcinogens 73 2. Nongenotoxic carcinngpne 75 C. The Role of Oncogenes and Suppressor Genes in Carcim o ’79 1. Retroviral oncogenesis 30 a. The acutely transforming retroviruses 30 b. The latent transforming retroviruses 34 2. DNA Tumor Viruses 36 a. Oncogenic Papovavirnsns 36 b. Oncogenic Adenovirnses 40 C. Oncogenic Herpesvirucec 41 d. DNA tumor viruses without oncogenes ..................... 42 3. Cellular Oncogenes in Carcin- o ' 44 a. Activation by point mutation 45 b. Activation by gene rearrangement 46 c. Activation through gene amplification 48 4. Oncogenes that Act as Suppressors to Neoplastic Growth ................. 50 a. The retinoblastoma paradigm an b. Other possible tumor suppressor genes 52 D. Oncogene Protein Products and their Role in Carcinogenesis ..................... 54 1. Extracellular oncogene proteins 55 2. Oncogene protein products as growth factor receptors ........................ 56 3. Oncogene proteins with phosphokinase activity ................... 58 4. Oncogene proteins located within the nucleus E. The ras Genes and Their Protein Products 63 1. The molecular biology of the ras genes 63 2. The ras protein structure 67 3. Posttranslational modification of the ras proteins 71 4. The biochemical activity of the ras proteins 73 F. Chemical Carcinogenic Risk A uent 76 1. Methods used in risk ‘ 76 a.) Epidemiological analysis in risk assessment ............................ 77 b.) Short-term tests for the evaluation of chemical carcinogenic potential 80 c.) The animal bioassay 84 i.) Basic conduct 95 ii) Problems with interpretation 85 G. Oncogene Activation in B6C3F1 Mouse Liver Tumors 90 1.) Activated oncogenes in spontaneous liver tumors 90 2.) Activation of cellular oncogenes in chemically induced 36C3F1 mouse liver tumors 94 Reference: 101 CHAPTER II. MUTATIONAL ANALYSIS OF THE H-ras ONCOCENE IN SPONTANEOUS 35C3F1 MOUSE LIVER TUMORS AND TUMORS INDUCED WITH A GENOTOXIC AND NONGENOTOXIC HEPATOCARCINOGENS ........................... 12 ABSTRACT 12% INTRODUCTION 124 MATERIALS AND METHODS 127 Tumor Induction 127 Oligonucleotide Primer Preparation. 127 PCR Amplification. 128 Sequencing 129 RESULTS 130 Tumor Induction 13o H-ras Gene Activation 1 '31 DISCUSSION 133 ACKNOWLEDGEMENTS 140 REFERENCES 141 CHAPTER III. A MODEL FOR 85C3F1 MOUSE LIVER TUMOR DEVELOPMENT: IMPLICATIONS FOR RISK ASSESSMENT 1 '53 APPENDIX 162 LIST OF TABLES W TABLE 1. Factors effecting cancer incidences 7 TABLE 2. Selected oncogenes: function and cellular localization ...................... 33 TABLE 3. Transforming genes in spontaneous hepatocellular tumors of the B(,C3F1 mouse 91 TABLE 4. Oncogene activation in B6C3F1 mouse liver tumors ......................... 93 TABLE 5. Mutation spectrum in codon 61 (CAA) of the H-ras gene in B(,C3F1 mouse liver tumors 9:; TABLE 6. Distribution of activating mutations within the H-ras gene from 86C3F1 mouse liver tumors 98 W TABLE 1. Incidence of Spontaneous and chemically induced liver tumors in male 136C3F1 mice 149 TABLE 2. Frequency of activation of the H-ras gene at codon 61 in male 36C3F1 mouse liver tumors 1'50 TABLE 3. Frequency of H-ras gene activation in hepatocellular adenomas and carcinomas in male B6C3F1 mice 151 TABLE 4. Mutation spectrum in codon 61 (CAA) of the H-ras gene in male B6C3F1 mouse liver tumors 1‘52 LIST OF FIGURES W FIGURE 1. Schematic representation of the carcinogenic process ........................ 8 FIGURE 2. Structural and functional domains of the mammalian ras gene...69 W FIGURE 1. Nucleotide sequence of the coding strand for the H-ras gene which was PCR amplified and sequenced 147 FIGURE 2. Representative DNA sequence of the H-ras gene codon 61 region derived from PCR amplified liver tumor DNA .............................. 148 W FIGURE 1. Representative chromatogram of HPLC purification for DNA oligonucleotide primers 166 FIGURE 2. Agarose electrophoretic gel analysis demonstrating the utilility of Bam H1 digestion on increasing PCR amplification efficiency ....168 FIGURE 3. Agarose electrophoretic gel analysis from a representative PCR amplification of a 110 bp region within the H-ras gene from liver tumor tissue 169 FIGURE 4. Representative DNA sequence analysis of the H-ras gene codon 12, 13 and codon 117 regions 170 vii INTRODUCTION Epidemiological studies suggest that the majority of human cancers are caused by environmental factors (Doll and Peto, 1981). Table 1 highlights a number of general factors responsible for the development of cancer in humans. A closer examination of this list reveals that these general factors represent variations of only three specific etiologic agents,- radiation, viruses and chemicals. Of these three, over 75% of induced human cancers are caused by the exposure to chemicals (the majority coming from natural sources). Because such a high percentage of human cancers are the result of chemical exposure, extensive efforts have been put forth by the scientific community to understand the mechanisms of chemical carcinogenesis and to develop experimental methodologies for the evaluation of a chemical's carcinogenic potential. Methodologies currently used to evaluate a chemicals carcinogenic potential include human epidemiological analysis, in vitro and in viva mutagenicity assays and the animal bioassay for carcinogenicity. Although data from all of these sources are used for chemical carcinogenic risk assessment, the results from the animal bioassay have the most significant influence on this decision making process. The development of tumors in these animals is not however definitive evidence that a chemical is a human carcinogen. There are a number of factors which complicate the extrapolation of bioassay results in determining potential human risk. Table 1. Factors effecting cancer incidences 1 Percent of all cancer deaths Factor or class of factors Best estimate Range of acceptable estimates Diet 35 10 - 70 Tobacco 30 25 -40 Reproductive and sexual behavior 7 1 - 13 Occupation 4 2 - 8 Alcohol 3 2 - 4 Geophysical factors 3 2 — 4 Pollution 2 < 1 - 5 Medicines and medical procedures 1 0.5 - 2 Food additives < 1 -53 _ 2 Industrial products <1 < 1 - 2 Infection 10 ? 1 - ? 1 From Doll and Peto, 1981 These complicating factors include such issues as differences in metabolism and pharmacodynamics between the test species and man, inherent problems with the bioassay design, and the occurrence of tissue specific high spontaneous tumor incidences within the inbred rodent strains. The issue of high spontaneous tumor incidences has been of particular concern with the B§C3F1 mouse. This mouse is used extensively in many carcinogenicity bioassay programs, including those conducted by the National Toxicology Program. The male 86C3F1 mouse has a characteristically high spontaneous liver tumor incidence. Increases in B6C3F1 mouse hepatic tumors are frequently observed in many bioassays after administration of either genotoxic or nongenotoxic chemicals. Evaluating the significance of an increase in these liver tumors to potential human risk after exposure to nongenotoxic agents has been a very controversial issue of the current bioassay program (Interdisciplinary Panel on Carcinogenicity, 1984; Nutrition Foundation, 1983; Task Force of Past Presidents- SOT, 1982). Greater confidence in the interpretation of bioassay results using this strain of mouse requires a better understanding of how these spontaneous tumors arise and how various chemicals enhance this tumor incidence. Efforts to understand the molecular mechanisms of carcinogenesis have been aided significantly by the recent detection of activated cellular oncogenes (Shilo et al., 1981). These genes appear to be involved in the control of cellular proliferation and of differentiation. These oncogenes have been found in a number of human and experimentally induced animal tumors. Recently, activated oncogenes have been identified in spontaneous liver tumors of the B(,C3F1 mouse and in tumors induced with several genotoxic chemicals (Fox et al., 1987; Reynolds et al., 1986, 1987; Wiseman et al., 1986). The most frequently detected oncogene in these tumors was the activated H-ras gene. Mutational analysis of a limited number of spontaneous tumors and tumors induced with several genotoxic carcinogens has revealed specific mutational spectra within the H-ras gene. One important parameter in evaluating a chemicals' carcinogenic potential is determining whether the chemical has in vivo genotoxic activity. The existence of mutational specificity in the H-ras gene suggests that a comparitive analysis between spontaneous tumors and tumors induced with various carcinogenic agents might be useful for making this type of evaluation. Such analysis would be particularly valuable for determining whether nongenotoxic hepatocarcinogens have any cryptic in vivo genotoxic activity since the current assignment of a nongenotoxic status is based primarily on in vitro mutagenicity assays. It may also provide important mechanistic information into how this class of carcinogens cause liver tumors in the B6C3F1 mouse. Information of this type could be very useful in the interpretation of bioassay results with this strain of mouse. This ultimately should provide for a more accurate evaluation of potential human risk from exposure to these chemicals. To address these issues, the objectives of this thesis are to (1) provide additional data on the frequency of H-ras gene activation in spontaneous B(,C3F1 mouse liver tumors (2) evaluate the mutational spectra of the H-ras gene in spontaneous liver tumors (3) determine the frequency of H-ras gene activation in tumors induced with several functionally diverse nongenotoxic hepatocarcinogens and (4) evaluate the mutational spectra in the hepatic tumors induced with these nongenotoxic agents. Chapter I provides a review of the scientific literature relevant to the proposed objectives. It includes discussions of the general process of carcinogenesis, chemical carcinogenesis, oncogenes and suppressor genes, a comprehensive review of the ras genes, chemical carcinogenic risk assessment and oncogene activation in the BGC3F1 mouse liver. Chapter II consists of an original research paper excepted for publication in the journal Cancer Research. Within this paper the objectives of the current thesis are addressed. The design and execution of the experimental work, including the development of a modification in the polymerase chain reaction (PCR) technique for amplifying designated regions of the H-ras gene from liver genomic DNA, presented in this paper was conducted by myself with the guidance of Dr. I. Justin McCormick. Histopathological evaluation of the liver tumors was conducted by Dr. Barry Yano. Instructional assistance on the synthesis of DNA oligonucleotide primers and modification in the DNA sequencing of PCR amplified H-ras genes was provided by Dr. Peggy Schott and Dr. Iia-ling Yang respectively. Chapter III presents a discussion of a theoretical model for the development of liver tumors in the 86C3F1 mouse. An Appendix has been included to provide additional information regarding specific methodological material that was not discussed in the Cancer Research manuscript. W AIheEmcessnLCarsinngenesis Cancer is a complex, dynamic disease which has been observed to occur in nearly every tissue of the body. It is a disease characterized by abnormal cellular proliferation and differentiation. These perturbations of growth and development give rise to a tumor (except in cancers of the reticuloendothelial system) which is the end result of a series of cellular changes that may have taken many years to develop. Although these changes occur as a continuous event, carcinogenesis can functionally be divided into three distinct stages, designated initiation, promotion and progression (Figure 1). Much of our understanding about the events occurring during these stages of carcinogenesis has been obtained from the development of experimental animal models using radiation and chemicals to induce specific cancers (Pitot, 1986). 1. I 'l' l' Initiation is considered to be a primary, irreversible change which predisposes the cell to further neoplastic development (Boutwell, 1964). The most likely site for this irreversible event is the genetic material (DNA) where the induction of mutations is thought to be a key step. This conclusion is based on the large amount of data associating the induction of mutations with the development of cancer. 8:238 88:80.22 . 32898. 88:88.8 . $9.: 0588.an . maotgafiogafi . 30.299: .mccoc . 29992: . c.9955. . =8 29% . 5:205 SEE . :meoaxo econ 0923 . engage: . c9.mu__nmficv 260.com . coacaaxo =8 . 89:3 <20 . \ 553565 T 228.55 T cozaEc. / «45> 8E3 E2982 :8 .232 Arlcozmfimm ”$2820 $305 2:32.830 05 Lo 5:25.332 czancom ._. 959”. Experimental evidence indicates that these initiating mutations may occur spontaneously or they can be induced by exposure to chemicals (naturally occurring and synthetic), radiation, and viruses. For many years the genetic target for these mutational events was unknown. However, over the past ten years considerable data have been generated implicating two groups of cellular genes as the likely targets. These two groups are represented by a set of dominant and recessive genes referred to as oncogenes and suppressor genes respectively. Druckey (1967) used the initiating carcinogen 4-dimethylaminostilbene to investigate the relationship between dose of initiator and tumor development in the rat. He observed a linear relationship between these two parameters with the line passing thru the origin. This study, along with data from many other studies where chemicals and radiation have been used as initiating agents, suggests that initiation is a "one hit, no threshold" process (Smets, 1980). Consequently, multiple divided doses of the initiator achieves the same result as a comparable single dose. Initiation alone is not sufficient for tumor development. The initiated cell may remain dormant indefinitely or grow very slowly. These "transformed" cells remain latent until they are acted on by some promoting event. 2. Protection Promotion is a complex process thought to involve several steps, some irreversible and some reversible. Experimental evidence indicates that 10 promotion does not involve mutational changes in the genetic material (Pitot, 1980). Instead promotion involves an increase in cellular proliferation and the alteration of gene expression (Angel et al., 1986; Krieg et al., 1988). The end result of this activity is the expansion of the initiated cell population. Cellular proliferation appears to be a necessary but not sufficient aspect of promotion. The requirement for cell division probably involves the fixation of the initiating event. The sooner the proliferative stimulus occurs after initiation the greater the probability that a neoplasm will develop. This is a reasonable observation if initiation is a mutagenic process because a delay of cell division would allow this damage to be repaired eliminating the initiating event. Promotional proliferation has been experimentally induced using both physical and chemical methods. In the rodent liver the rapid regeneration following surgical partial hepatectomy has been shown to induce tumors in the chemically initiated rat liver where chemical treatment or partial hepatectomy alone will not cause tumor development ( Pitot, 1980a). Chemical agents used to promote rodent skin and liver tumors also induce cell division (O'Connell et al., 1987). The increased cellular proliferation by these agents may be the result of a degenerative-regenerative response due to the cytotoxic effects of the chemical or to a specific pharmacologic stimulus inducing the activation of genes involved in cellular division. These agents include a diverse group of materials such as hormones, drugs, plant products and industrial chemicals. 11 Chemical tumor promoters can be divided into specific and nonspecific classes. The specific promoters alter gene expression by forming complexes with unique receptors or receptor-like molecules. These receptor-ligand complexes can activate a series of biochemical reactions leading to changes in transcriptional activity. Examples of this class of promoter include steroid or polypeptide hormones (Pitot et al., 1980a) and the potent tumor promoter TCDD (Pitot et al., 1980b; Poland et al., 1975). Nonspecific promoters do not interact with specific receptors but induce changes in gene expression by indirect mechanisms that are not clearly understood. Experimental studies have demonstrated that chemical promoters, like chemical initiators, exhibit a dose response relationship. However unlike initiating agents, tumor promoters appear to have a distinct threshold concentration below which tumors will not occur and a maximum concentration where a further increase in tumor development is not observed. This concept has been demonstrated using either a mouse skin or rat liver model with TPA or phenobarbital as the promoting agent respectively (Verma and Boutwell, 1980; Peraino et al., 1980). Pitot (1982) and his associates, using quantitative measurements of enzyme altered foci as a parameter to study chemical promotion in the liver, showed that a maximum number of foci occurred at doses above 0.01% phenobarbital in the diet. At the low dose of 0.0001% no difference in the number of foci was detected between the treatment and the control groups. Studies such as these clearly exemplify the distinct differences between chemical initiation and promotion. This concept is not only important to an academic understanding 12 of the process of carcinogenesis but also has practical implications for chemical risk assessment. 3- mm After the promotional phase of neoplastic development the growing tumor consists primarily of a clone of cells which morphologically are very similar to the surrounding normal cells. These early tumors do however exhibit biochemical differences that can be exploited to distinguish them from the surrounding background of normal cells. Using various histological staining techniques they can be identified microscopically as areas of altered focal development (Pitot, 1980). These developing tumors are pathologically classified as benign. During the period of progression these benign tumors become more malignant in character. This involves a process of tumor evolution where the sequential selection of variant sub-populations occurs which exhibit progressive phenotypic changes that are of a more neoplastic character. One of these characteristics is a tendency for the developing tumor to exhibit an increasing growth rate. This does not represent a shortening of the cell cycle but rather involves an increase in the proportion of cells which are actively dividing instead of progressing towards terminal differentiation (Ford, 1986). In some cases this release from normal growth control may involve an altered response to growth factors through the loss of specific receptors or by an autocrine stimulation of growth caused by the unregulated overproduction of a growth factor (Lippman, 1986). 13 As the tumor becomes more malignant a decrease or disappearance of cellular organelles and metabolic functions necessary for specialized activity is observed within cells of the evolving tumor (Croce, 1985). At the same time significant increases in the synthesis of specific proteins may occur which contributes to the ability of the tumor to invade or metastasize. These proteins can include proteolytic enzymes, plasminogen activators, fibronectin, and tumor angiogenic factors (Nicolson, 1984). Decreased antigenicity and the acquisition of drug resistance may also appear within the advanced neoplastic cell population which could provide those cells with a selective growth advantage (Nowell,1986). One of the more pronounced changes occuring during progression is an increase in the frequency of cells with chromosomal abnormalities. Histological studies of isolated tumor masses have demonstrated that tumor cell populations become more genetically unstable with increasing time (Oksala et al., 1979; Ling et al., 1985). These cells show higher numbers of chromosomal breaks, nondisjunctions, increased sister chromatid exchanges and changes in ploidy. Some limited data also suggest that the rate of these chromosomal changes increases with tumor progression (Ling et al., 1985). These chromosomal changes are probably not occuring randomly. Evidence has been accumulating suggesting that specific chromosomal changes are associated with particular malignancies. The most striking examples of this can be found in several human leukemias. For example in Burkitt's lymphoma, an aggressive Bocell neoplasm, approximately 75% of the patients exhibit a chromosomal translocation in which the distal part of 14 chromosome 8 is translocated to the long arm of chromosome 14 (Manolova and Manolova, 1972). Variant translocations involving chromosome 8 with chromosome 22 and chromosome 2 are found in the remaining 25% of these patients (16% and 9% respectively) (van den Berghe et al., 1986). Most cases of chronic myelogenous leukemia (CML) involve a translocation between chromosome 9 and chromosome 22 producing a marker chromosome which has been named the Philadelphia chromosome (Nowell, 1960). Recently the myc and abl oncogenes have been located on the fragments of chromosomes 8 and 9 involved in the translocations of Burkitt's lymphoma and CML respectively (Dalla Favera, 1982; Groffen, 1984). Alterations in the expression or in the protein coding region of these genes as a result of the translocations is thought to play an important function in generating these specific malignancies. Single gene mutations could initiate the chromosomal aberrations observed during tumor progression. Defects in DNA repair enzymes or abnormalities resulting in more error prone DNA synthesis or mutations in mitotic spindle proteins could precipitate some of the changes described above. Once a single chromosomal abnormality occurs it in itself may contribute to further genomic instabilities. As cells in the developing tumor begin to accumulate increasing numbers of genetic changes, eventually a cell(s) may emerge which has acquired a completely malignant phenotype. As this cell proliferates its progeny begin to invade the surrounding tissue. Some of these cells may find their way into the vascular or lymphatic systems where they can be 15 disseminated throughout the body giving rise to metastatic tumors leading to the ultimate death of the organism. The multi-stage cancer model of initiation, promotion and progression serves as a paradigm to recognize the functional importance of the major critical events in tumorigenesis. This model depicts tumor development as a series» of distinct steps occuring temporally one after the other. It is important to remember that the formation of a tumor from a single initiated cell to the appearance of a malignant mass is a continuous process representing many biological changes that may overlap each other. In spite of these limitations, the multi-stage cancer model has made an important contribution in developing our current understanding of the cancer process. 4. W The classification of a tumor as benign or malignant is based on a subjective criterion incorporating specific growth and morphological characteristics. If one uses these criteria most tumors fall into one of these two categories. However, there are some borderline cases which cannot be categorized with certainty. All tumors, benign or malignant, have two basic things in common; (1) they consist of proliferating neoplastic cells of parenchymal origin and (2) have a supporting stroma consisting of connective tissue, blood vessels and possibly lymphatics (Cotran et al., 1989). Beside these two common parameters, differences distinguishing benign and malignant tumors exist in the level of differentiation, rate of growth and type of growth. 16 In general, cells comprising a benign tumor are well differentiated. Cytologically the tumor cells do not differ substantially from the surrounding normal cells within the organ. In many cases they may present themselves as outgrowths of the normal tissue. For example, bone tumors may produce nodules that are indistinguishable from normal bone. Benign epithelial tumors of the skin form local growths which contain all of the components of normal skin but are compacted to form a solid nodule (Franks, 1986). An example of this type of tumor is the common wart. As the growing tumor expands it may compress and damage the normal tissue. This can lead to the severe loss of organ function. Depending on the particular organ and the extent of damage some benign tumors, although rarely, can be fatal. On the other hand malignant tumors are composed primarily of undifferentiated cells. It is thought that they may originate from the pool of undifferentiated stem cells found in all specialized tissues (Cotran et al., 1989). The cells within the tumor characteristically display morphological pleornorphisms. Cells much larger and much smaller than normal can be found. These cells can be multi—nucleated or contain a single giant polymorphic nucleus. Large numbers of cells in mitosis are frequently observed indicative of the increased proliferative activity associated with malignancy. In general, the malignant tumor shows little resemblance to the surrounding tissue. The growth characteristics of the cells within a tumor are another important parameter used to distinguish a benign tumor from malignant one. Most benign tumors grow slowly over a period of years at a steady pace. 17 They may however enter periods of dormancy where they do not grow at all or even decrease in size. Factors such as hormone dependency and blood supply can influence this growth rate ( Cotran et al., 1989). Growth of a benign tumor is nearly always confined within a fibrous capsule. This contained, localized growth is a key factor distinguishing benign from malignant neoplasms. However, the lack of a capsule does not imply that the tumor is malignant. The growth rate of cells within a malignant tumor is usually rapid and sometimes erratic. The pattern of growth is much more unorganized than that found in a benign tumor. Malignant tumors have the capacity to invade the nearby surrounding tissue, although carcinomas in situ can still be diagnosed as malignant based on cellular characteristics. The unequivocal diagnosis of malignancy is the detection of a metastasis. A metastasis is a secondary tumor arising from the primary tumor at some other location in the body. Metastases can occur via several different mechanisms (1) embolization through the blood vessels (2) lymphatic permeation and (3) direct seeding of the body through open cavities (e.g. peritoneal cavity) (Cotran et al., 1989)). The properties necessary for tumor cells to metastasize are most likely acquired during the period of tumor progression. 18 5. Wasted: The ability of a tumor to invade and metastasize ultimately results in the death of the cancer patient. It makes treatment difficult and if metastasis could be eliminated most cancers would be curable by surgery. Tumor metastasis is a complex, dynamic process involving a number of specific sequential events resulting in the dissemination of the primary tumor to various organs around the body. Following tumor growth and development, usually during the period of progression, metastasis begins with the local invasion and infiltration of tumor cells into the surrounding normal tissue (Cotran et al., 1989). Some of these cells penetrate the small vessels of the lymphatic or vascular system. For a metastatic cancer cell to enter the lymphatic or vascular vessels it must first traverse the extracellular matrix (ECM). This matrix is composed of a highly cross-linked meshwork of type IV collagen, specific glycoproteins such as laminin, and several types of proteoglycans (Cotran et al., 1989). Studies have shown that the invading cells actively secrete enzymes which breakdown the ECM components (Pauli and Lee, 1988). Once the tumor cells have penetrated the vasculature, individual cells or clumps of cells are released into the circulation (Kaminski et al., 1988). The circulatory system is a hostile environment for the cancer cell and many cells do not survive. This is the result of several factors which include unfavorable growth conditions and the action of immunological factors which tend to control the spread of the tumor cells. Once into the circulation the tumor cells tend to aggregate with themselves and with platelets to form 19 small emboli. The aggregation of the tumor cells with the platelets may increase survival by protecting them from immunological attack. Some of the circulating emboli eventually come to rest within the capillary beds of various organs. Here they exit the vasculature and begin growing as a secondary tumor completing the metastatic process. Studies have shown that some primary tumors form metastases within specific tissues (Liotta, 1988). It is not entirely clear what factors influence where these secondary tumors will arise. This selectivity may in part be a function of the anatomical location of the primary tumor. The likelihood of whether circulating tumor cells find their way to a particular organ may also be influenced by the patterns of blood flow from the primary tumor site. These natural patterns of blood flow cannot entirely explain metastatic specificity. This is exemplified by the fact that muscle tissue and the kidneys are both well vascularized yet these organs seldom are the sites for metastases (Cotran et al., 1989). Cell surface factors, most likely involving common membrane receptors between the tumor cells and the cells where the metastases occur, probably play a more important role (Pauli and Lee, 1988). l 6. W As discussed previously, the stage of tumor progression is characterized as a time when the cells of the developing tumor become increasingly heterogeneous. A number of sub-populations will arise within the developing tumor which have their own characteristic phenotype. This 20 heterogeneity involves an extensive number of phenotypic changes including cellular morphology, tumor histology, karyotype, growth rate, cell products, receptors, immunological characteristics, metastatic ability and sensitivity to therapeutic agents (Heppner, 1984). Much of the experimental work attempting to explain the source of this heterogeneity has focused on the inherent genetic instability of the transformed cell. The work of Cifone and Fidler (1981) demonstrated that in fibrosarcoma cells the rate of spontaneous mutation was greater in metastatic cells than in nonmetastatic subpopulations. Mutations such as genomic rearrangements, chromosomal losses, gene amplification and base pair substitutions are involved in the increased phenotypic varibility. It is important to recognize that heterogeneity is not a process restricted to tumor cells. Phenotypic variability also occurs in normal tissues. Mechanisms generating tumor heterogeneity may therefore be similar to those involved in the development of normal tissues. This is exemplified in the observation that normal cells differ in their susceptibility to various carcinogens and oncogenic viruses. For example, the heterogeneity observed in different SV40-transformed clones reflects the inherent heterogeneity associated within the normal cell population prior to transformation (Kakunaga, 1980; Omar, 1983). . This normal variability probably arises as a consequence of epigenetic events involving changes in gene expression. Evidence implicating the importance of epigenetic processes is provided by Harris et al. (1982) and Peterson et al. (1983) where they reported the rates of variation / generation to 21 be 10'2 t010'5 which is several orders of magnitude greater than the rates of mutation (10'6 to 10'8). Unequivocal proof that tumor heterogeneity is a real phenomenon of tumorigenesis and not an artifactual observation is still unavailable. The concept of heterogeneity is however generally accepted by most pathologists and minor biologists. One corollary of the tumor heterogeneity hypothesis, which is far more controversial, states that metastases are derived from one or more of the heterogeneous sub-populations within a single primary tumor. Recent work by Kerbal and his associates has provided convincing evidence that metastases are derived from variant sub-populations within a single tumor (Wanghorne et al., 1988). These investigators generated a number of transformed clones from single tumors each carrying a selectable marker on a different DNA restriction fragment. Injecting these cells into animals they were able to generate primary tumors that eventually gave rise to tumor metastases. By comparing the restriction fragments of the marker gene in the primary and metastatic outgrowths they could trace the cell lineage of the secondary tumors. Using this approach they demonstrate that the metastatic tumors carried the marker gene on the same restriction fragment as was found in the primary tumor. Unexpectedly, it was also observed that the metastatic cells eventually achieved a dominant status within the heterogeneous cell population of the developing tumor. EQemicaLCm'mcgrnesis One of the earliest indications that chemicals can be carcinogenic comes from the now classical studies of Percival Potts, the eighteenth century physician, who made the astute observation that the increased incidence of scrotal cancer in chimney sweeps was associated with exposure to soot. Early in this century two Japanese workers demonstrated that coal tar in soot could induce tumors when directly applied to the skin of rabbits (Yamagiwa and Ichikawa, 1918). Later Kennaway chemically fractionated coal tar and showed that the pure components dibenz[a,h]anthracene and benzo[a]pyrene had carcinogenic potential (Williams and Weisburger, 1986). During the late 1800's the German physician Rehn observed that workers in the dye industry had an abnormally high incidence of bladder cancers. It was not, however, until nearly 50 years later that Hueper demonstrated that 2—naphthylamine was the causative agent by reproducing these lesions in dogs (Williams and Weisburger, 1986). Since these early observations several hundred chemicals have been identified as having carcinogenic activity. Many of these carcinogens are synthetic and include the general classes of chemicals such as the polycyclic aromatic hydrocarbons, heterocyclic amines, nitro alkyls, halogenated hydrocarbons, aldehydes, carbamates, and N-nitroso compounds (Williams and Weisburger, 1986). In addition to these synthetic chemicals, a number of natural materials are potent human and/ or animal carcinogens. These carcinogens can be found occuring naturally in our food or as a result of cooking. For example the plant product cycasin found in certain foods eaten by peoples in the Pacific islands was found to have potent carcinogenic activity (Williams and Weisburger, 1986). Raw mushrooms contain various hydrazines and certain bracken ferns have been shown to cause cancer of the alimentary tract (Wigley, 1986). Other naturally occurring carcinogens can be formed as a result of the metabolizing activity of micro-organisms. Aflatoxin Bl, a mycotoxin produced by a strain of Aspergillus flavus, is one of the most potent known carcinogens and has been implicated in the high incidence of liver cancer present in many third world countries. (Williams and Weisburger, 1986). Once sufficient evidence had been accumulated implicating chemicals as an important etiologic agent of cancer causation efforts were made to understand the mechanisms of chemical carcinogenesis. From these efforts chemical carcinogens are now grouped into two general categories based on their mechanism of action. These categories are referred to as genotoxic and nongenotoxic carcinogens. 1. Warns Genotoxic carcinogens are defined as having the ability to react with and cause damage to the DNA. By definition these agents are also mutagens. This class of carcinogens consists mainly of materials that are chemically electrophilic. A few of the genotoxic carcinogens, for example the alkylating 24 and acylating agents dimethyl sulfate, ethyl nitrosourea, dimethylnitrosamine and sodium azide, are electrophilic enough to react directly with the DNA (Mch et al., 1975). However, for the majority of chemical carcinogens metabolic activation is required to convert the parent material to a more reactive electrophilic species (Miller and Miller, 1981). This metabolic conversion is accomplished primarily by a group of enzymes located in the microsomal fraction of the endoplasmic reticulum referred to as the cytochrome P-450 dependent mono-oxygenases. Electrophilic activation may occur as a single step reaction or may involve multiple steps. At the same time other reactions may be taking place to detoxify the parent material or its metabolic intermediates. Therefore the potency of any particular genotoxic carcinogen involves the balance between these competing reactions. Regardless of the particular mode of activation the end result is the production of an electrophilic molecule that can react with DNA. No single unique alteration of the DNA is responsible for the carcinogenic action of a chemical carcinogen/mutagen. Chemical interaction with the DNA can produce the formation of specific DNA base adducts that can lead to base pair substitutions. Some chemicals can intercalate between the bases causing frameshift mutations, additions or deletions. Other carcinogens induce gross alterations of the genetic material producing chromosomal abnormalities such as translocations or chromosomal breaks. All of these mutagenic changes can have carcinogenic consequences. 1W Not all carcinogenic chemicals have mutagenic activity. There is an important class of chemical carcinogens, generally referred to as epigenetic or nongenotoxic carcinogens, which do not interact directly with DNA but still can cause the formation of tumors in animals and humans. This class of carcinogens is composed of a diverse group of pharmacologically and toxicologically distinct chemicals and materials. Examples include certain chelating agents, hormones, pharmaceutical drugs, solid state materials (plastics and metals) and various synthetic organic chemicals. Little is known about the specific mechanisms by which these agents cause tumor development. It has been suggested that these agents may exert their carcinogenic activity by promoting previously initiated cells or through the induction of somatic mutations by indirectly damaging DNA as a consequence to some abnormally induced physiologic response (Schulte- Hermann et al., 1983). Studies conducted a number of years ago may have provided a possible explaination as to how nongenotoxic agents can be carcinogenic. These early studies demonstrated that chronic tissue irritation can cause the development of tumors in rodents at the site of the irritation (Berenblum, 1929). This irritation can be brought about through repeated freezing or by repeated subcutaneous injections of nonreactive materials such as glucose, saline, and distilled water. This continuous irritation can cause repetitive tissue damage resulting in an increase in focal cellular proliferation. It is believed that the 26 increasein cellular proliferation is associated with the development of these tumors. Increased cellular proliferation is also frequently observed at tumorigenic doses for many animal nongenotoxic chemical carcinogens. This is particularly true for a number of small chlorinated hydrocarbons such as chloroform, tetrachloroethylene and trichloroethylene (Loury et al., 1987) This increase in cellular division appears to be a consequence of a toxicity induceddegenerative/ regenerative response within the target organ. It is not entirely clear how increased cellular proliferation can result in the enhancement of tumor development. It has been well established that mutations occur during normal DNA synthesis. As reviewed by Loeb et al. (1978) and Hartman (1980), base mismatches may arise from polymerase base selection errors or as a result of induced errors by proof reading enzymes. Estimates of in viva somatic cell mutation rates vary from 10'9 to 10'12 events per nucleotide per cell division (Drake et al., 1969). It stands to reason that an increase in the number of replicating cells will result in an increase in the mutation frequency. This in turn could lead to an increase in the number of cells carrying mutations within important cancer genes (oncogenes or suppressor genes). The end result could be an increase in the population of initiated cells or the induction of additional mutational changes required for the further neoplastic progression of previously initiated cells. Either one of these mechanisms could lead to an increase in tumor development. 'As was discussed earlier, an important component of tumor promotion is the expansion of the initiated cell population. Increases in 27 cellular proliferation as a result of exposure to some nongenotoxic agents may increase tumor development by promoting previously initiated cells. Thus, nongenotoxic compounds could increase tumor yield without causing any de novo initiation. Other nongenotoxic carcinogens do not appear to exert their tumorigenic action through increases in cellular proliferation. Alterations in gene transcription, perturbations in signal transduction pathways, inhibition of intercellular communication and peroxisome proliferation have also been proposed as mechanisms of nongenotoxic carcinogenicity (Butterworth and Slaga, 1987). For example, the very potent rodent hepatocarcinogen tetrachlorodibenzo-p—dioxin (TCDD) has a specific cytosolic receptor protein with a high affinity for TCDD. This protein / TCDD complex translocates to the nucleus of the liver cell and increases the transcriptional activity of certain genes (Poland and Knutson, 1982). Tetradecanoyl phorbol acetate (TPA) has been classically considered a tumor promoter based on the two-stage mouse skin model. However TPA by itself can cause a low incidence of skin tumors and when applied to hairless mice elicits a 100% incidence of skin cancer (Inversen and Inversen, 1979). These carcinogenic effects may in part be due to the ability of TPA to bind directly to and activate protein kinase C. This inappropriate activation of protein kinase C results in the perturbation of signal transduction pathways altering gene expression which could have transforming consequences. Some of these nongenotoxic agents may exert their carcinogenic effect by causing indirect damage to the DNA. It has been known for some time that many tumor promoters can activate polymorphonuclear leukocytes (PMNs) and that this activation leads to the formation of large amounts of active oxygen radicals (Goldstein et al., 1981). When these cells are activated by nonphagocytic stimuli such as tumor promoters, these oxygen radicals are free to react with neighboring cells and can damage cellular macromolecules such as DNA. Several investigators have demonstrated that these PMN generated oxygen species can induce DNA strand breaks and oxidized bases such as 5-hydroxymethyl-2'-deoxyuridine and thymidine glycol in coincubated cells (Dutton and Bowden, 1985; Frenkel and Chrzan, 1987). A similar mechanism of action may be involved with another class of nongenotoxic animal liver tumorigens which are commonly referred to as peroxisome proliferators. Some of these peroxisome proliferators include the hypolipodemic drugs clofibrate, ciprofibrate, fibric acid, fenofibrate, the plasticizer di(2-ethylhexyl)phthalate and the organic solvent 1,1,2 trichloroethylene (Elcombe et al., 1984; Reddy et al., 1983; Kluwe et al., 1985). A common physiologic response to exposure from these chemicals is a large increase in the production of intracellular peroxisomes (Reddy et al., 1980). These cytoplasmic organelles are involved in lipid metabolism. One consequence of lipid metabolism is the production of H202 which is a highly reactive molecule. It has been proposed that the increase in the number of peroxisomes results in an excess production of hydrogen peroxide that may escape normal detoxification mechanisms and react with cellular lipids to initiate the auto-oxidation of fatty acids. Some of the products of this fatty acid metabolism include a number of reactive oxygen species (ie. aldehydes, 29 organoperoxides and conjugated dienes) that could react and damage DNA (Reddy and Lalwani, 1983). Although the mechanisms of nongenotoxic carcinogenesis appear to be many and diverse, the one obvious similarity is that they all seem to exert their carcinogenic activity at dose levels which are high enough to elicit some abnormal physiologic response (ie. toxicity, peroxisome proliferation etc). This has important implications in the evaluation of chemical carcinogenic risk for humans. It suggests that a carcinogenic threshold exists for these compounds which implies that if the exposure is low enough so that the physiologic response is not induced, then the development of tumors should not occur. C. W In the late 1960's Huebner and Todaro (1969) proposed that specific genes (which they named oncogenes) are involved in the genesis of cancer. Support for the existence of these genes has been provided by research demonstrating that (1) cells can be transformed in vitro or tumors induced in viva by exposure to radiation and mutagenic chemicals, (2) damage to DNA is associated with cellular transformation, (3) a hereditary prediposition exists for certain types of cancer, and (4) defects in certain DNA repair processes are associated with specific cancer predispositions. Since the conception of this hypothesis, cancer genes have been identified in a variety of mammalian cancer cells. The first oncogenes were 30 however originally identified in a specific class of animal tumor viruses, the retroviruses. ' 1. W The type-C retroviruses represent a group of RNA tumor viruses that can induce leukemias, lymphomas, sarcomas, and carcinomas in many different animal species including man (Bishop, 1983). The genome of these viruses consists of a single strand of RNA approximately 10 kb in length (Bister and Duesberg, 1983). Three genes, required for viral replication (gag, pal, env), are bound on each side by several hundred bases of untranslated, repetitive RNA referred to as long terminal repeats (LTR's). These LTR's contain transcriptional promoter and enhancer sequences in addition to a 5' cap and 3' poly -A tail. The gag gene codes for a structural protein that is associated with the RNA; the pol gene makes the reverse transcriptase used to synthesize a complementary DNA from the RNA; and the env gene codes for a glycoprotein which resides in the lipoprotein envelope of the viral particle. Based on their transforming properties these viruses are divided into 2 general groups referred to as (1) the acutely transforming retroviruses and (2) the latent transforming viruses. a. The acutely transforming retroviruses The acutely transforming viruses are characterized by the ability to induce rapid tumor development in infected animals. These viruses can produce tumors because they contain an oncogene within their viral genome 31 (Bister and Duesberg, 1982). All of the acutely transforming viruses (except the Rous sarcoma virus) have deletions in one or more of the three essential genes for replication. As a result these viruses are replication defective. Viral progeny can however be produced if these viruses are co-infected with non- transforming helper viruses. Once the virus enters the cell, the RNA is converted into DNA which integrates within the infected cell's genome. Expression of the viral oncogene results in the rapid transformation of the cell and eventual tumor development. The different transforming retroviruses are distinguished from each other by the particular oncogene residing within the virus. There are approximately 20 different retroviral oncogenes which can induce most of the major forms of neoplasms (Bishop, 1987). These viral oncogenes have evolved from normal cellular genes. They have been. acquired by the virus as a result of a viral transduction event (Bishop, 1983). During the evolution of the virus, specific structural changes (such as base substitutions, additions and deletions) have taken place within the normal cellular gene to produce the transforming properties of these viral oncogenes. The transforming capabilities of these viruses rests not only with the acquisition of an oncogene but with the ability to produce very high levels of the oncogene product. This capability for the high level of gene expression is due to the presence of the transcriptional enhancers located within the LTR's of the viral genome (Bister and Deusberg, 1982). A variety of structurally diverse proteins are encoded by the transforming retrovirus oncogenes. These genes can be grouped using 32 several different criteria. One that is particularly useful is based on location and biochemical activity of the protein products. Table 2 provides a partial list of some of the more common oncogene protein products. In general, the products and transforming function of the viral oncogene proteins are not well understood. The ability of these proteins to transform cells is undoubtedly related to their ability to disrupt the regulation of normal cellular proliferation and/ or differentiation. A more detailed discussion of the biological function of some of these oncogene protein products is provided in a later section. 33 Table 2. Selected oncogenes: function and cellular localizationa Oncogene Retrovirus Cellular Property or location function of protein Growth factors sis simian sarcoma extracellular related to PDGF ' . . virus int-2 - - extracellular related to fibroblast growth factor Protein kinases src Rous sarcoma plasma tyrosine kinase virus membrane yes avian sarcoma plasma tyrosine kinase virus membrane abl Abelson murine plasma tyrosine kinase leukemia virus membrane fes feline sarcoma plasma tyrosine kinase virus membrane fps feline sarcoma plasma tyrosine kinase virus membrane mos Moloney murine cytoplasm serine/ threonine sarcoma virus kinase raf avian sarcoma serine/threonine virus kinase Growth factor receptors erb-B avian transmembrane EGF receptor erythroblastosis tyrosine kinase neu - transmembrane similar to EGF receptor fms feline sarcoma transmembrane CSF-1 receptor tyrosine kinase ros avian sarcoma transmembrane structure related to growth factor Table 2 (cont) Oncogene Retrovirus Cellular Property or location function of protein GTP binding H-ras Harvey murine plasma GTP binding/GTPase sarcoma virus membrane binds with GAP K-ras Kirsten murine plasma GTP binding/GTPase sarcoma virus membrane binds with GAP N-ras - plasma GTP binding/GTPase membrane binds with GAP Nuclear myc avian nucleus binds d.s. DNA myelocytomatosis L-myc - nucleus binds d.s. DNA N-myc nucleus binds d.s. DNA ski avian SKV770 nucleus ? ' myb avian nucleus ? . myeloblastosis ets E26 nucleus ? fos murine nucleus AP-1 transcription osteosarcoma factor jun avian sarcoma nucleus AP-I transcription virus-17 factor a Derived from Cotran et al., 1989; Land et al., 1983; Hunter T., 1984. b. The latent transforming retroviruses These viruses are replication competent and have the typical retroviral genomic organization discussed above. They do not have transforming oncogenes, but can still cause the development of tumors after a long latency 35 period following infection. The mechanisms by which they cause neoplastic transformation is frequently referred to as "insertional mutagenesis" because it is the insertion of the proviral DNA at critical sites within the host genome that leads to cellular transformation (Bishop, 1987). If this insertion occurs within a cellular gene, structural modifications may take place leading to the oncogenic conversion of the gene. Alternatively, the presence of strong viral promoters near a cellular gene may lead to the production of high levels of the protein or result in the inappropriate expression of the gene. Thus, both genetic damage and / or altered gene expression may underly the neoplastic transforming properties of these viruses. Many of the latent transforming animal retroviruses are associated with the development of chronic leukemias. One human retrovirus, HTLV- 1, appears to be linked to the development of adult T-cell leukemia (ATL) in humans (Ehrlich and Poiesz, 1988). Evidence associating this virus with the etiology of ATL in humans includes the fact that (1) this virus has repeatedly been isolated from patients with ATL, (2) HTLV-l proviral sequences have been found in leukemic cells and not in normal cells from these patients, (3) cultured neoplastic cells release H'TLV-l that can immortalize T cells in vitro, and (4) antibodies against HTLV-l are found in over 90% of patients with ATL . The genomic structure of HTLV-I is typical of other latent transforming retroviruses in that it contains the 3 essential genes flanked by the LTR sequences. This virus differs in that it contains a fourth coding region located between the env gene and the 3' LTR which has been named tat (Nerenberg et al., 1988). The tat region codes for at least 3 proteins and it is believed that these proteins are partially responsible for the transforming activity of the virus. One protein causes an increase in HTLV-l transcription by acting on the viral 5' LTR. The other two proteins induce the activation of interleukin-2 (IL-2) and the IL-2 receptor. Thus, infected cells produce increased amounts of a growth factor and its receptor which would act as an autocrine stimulation for cellular proliferation. This process alone cannot cause cellular transformation indicating that additional changes must also take place, once again demonstrating the multi-step process of tumor development. 2. W Several DNA viruses have been associated with the development of cancer in animals and recent evidence suggests they are involved in some human cancers. These viruses belong to several different taxonomic families which include the Papovavirus, Adenovirus, Herpesvirus, and Hepadenovirus groups. a. Oncogenic Pamvaviruses Members of the Papovavirus group are of particular interest because they include two viruses, SV-40 and polyoma, of which more is probably known about their mechanism of transformation than about that of any other DNA tumor virus. In addition, this family includes the human papilloma virus (HPV) which has been associated with development of 37 several types of human cancers including in situ and invasive cancer of the cervix (Bonfiglio,1988). The genomes of both the SV-40 and polyoma viruses are composed of a small, circular, double strand of DNA approximately 5 kb in length (Dulbecco and Ginsberg, 1980). Infection by the virus is divided into two phases: an early phase .in which the viral DNA is replicated and a late phase during which viral proteins are synthesized and assembly of viral particles takes place. Each phase of the infection cycle is accompanied by a specific program of gene expression. These viruses can produce a productive or nonproductive infection depending on the species of the infected cell. Productive infection occurs in cells referred to as permissive. In these cells expression of both the early and late genes occurs resulting in the synthesis of large numbers of viruses which eventually kill the cell. Each type of papova virus has a narrow range of permissive cells in which expression of both early and late genes results in a lytic infection. For SV-40 and polyoma viruses, monkey and mouse cells respectively are the permissive hosts (Delbecco and Ginsberg, 1980). If infection occurs in a nonpermissive cell, viral DNA synthesis does not occur and the late genes are not expressed (Darnell et al., 1986). The expression of the early genes however still takes place which induces the cell to enter the S phase of the cell cycle. As long as the expression of the early viral genes occurs the cell is continually stimulated to proceed through the cell cycle. Most of the infected nonpermissive cells eventually revert to a normal phenotype as a result of the degradation or dilution of the viral 38 genome. If the viral DNA integrates into the genome of the host cell in a manner resulting in the constitutive expression of the early genes with the subsequent loss of late gene expression, then those cells become permanently transformed (Delbecco and Ginsberg, 1980). This integration involves a nonspecific type of recombination with the sites of integration occurring within the viral late gene region and occurring randomly within the recipient DNA. ' These viruses therefore owe their transforming capabilities to the proteins of the early genes. The early gene region of SV-40 codes for two proteins; large T and small t, and the polyoma early gene region codes for three proteins; large T, middle T, and small t. The early region of the polyoma genome is transcribed as one large mRN A molecule (3000 bp) that is spliced in three different ways to produce the appropriate transcripts for the three early proteins (Rassoulzadegan et al., 1982) Using molecular cloning techniques, plasmids have been constructed containing sequences coding for each of the early region polyoma proteins. Transfecting these plasmids into various cell lines has established that the polyoma large T protein can induce cellular immortalization whereas the middle T protein can transform previously immortalized cells (Gross, 1983). The mechanisms by which these proteins elicit their phenotypic response are not entirely understood. Recent advances have provided some important insights into the molecular mechanisms associated with the transforming properties of these proteins. The middle T protein of polyoma has been shown to associate with the cellular src protein at the inner aspect of the 39 cytoplasmic membrane (Courtneidge and Smith, 1983). This association increases the specific tyrosine kinase activity of the src protein by as much 50- fold. The middle T protein accomplishes this activation apparently by blocking the phOSphorylation of specific tyrosine residues on the c-src protein that when phosphorylated down regulates its biochemical activity. A similar process appears to be involved with the large T protein of the SV40 virus. Based on structural and genetic studies, the large T protein has two domains which are involved in the transformation of cells in culture (Srinivasan et al., 1989). One of these domains imparts the immortalizing function of the protein. This domain contains the binding site for the cellular protein p53 and forms a stable complex with this protein in cells infected and transformed with the virus (Lane and Crawford, 1979). The p53 protein has been associated with the process of neoplastic tranSformation in a number of cell types (Dippel et al.,1981; Oren et al.,1981). Recent studies by Finlay et al. (1989) support the idea that the p53 protein may act as a suppressor to block cell transformation. The transforming activity of the large T protein may in part be due to an inactivation of the suppressing activity of p53 through the formation of this protein complex. The second domain of the large T protein forms a complex with another known tumor suppressor gene, the retinoblastoma protein (p105-RB) (DeCaprio et al.,1988). Inactivation of this protein by the formation of the large T/ p105-RB complex is also thought to play an important role in the transforming activity of the SV40 virus. A similar complex is formed with the human papilloma virus -16 transforming protein E7 and the RB protein (Dyson et al.,1989). The formation of this complex is not surprising based on the structural similarity between the E7 protein and the SV40 large T antigen. b. Oanenic Adenoviruses Oncogenic members of the Adenovirus family include types 2, 5 and 12 (Delbecco and Ginsberg, 1980). None of these viruses cause tumors in humans but they can induce the development of sarcomas in the nonpermissive hamster or rat. The genes required for transformation are localized in the 6-7% extreme left region of the viral genome involved in early gene expression. As seen with the papova viruses, transformation occurs as a result of expression in the early genes. Two transforming genes, Ela and Elb, have been localized within this early region. Transformation of rodent cells in culture can be achieved if they are transfected with recombinant plasmids that contain only these early genes (Iochemsen et al.,1982). 4 The role of the Ela and Elb proteins in cellular transformation has been extensively examined. Ela has been shown to immortalize cells in culture when transfected with this gene (Houweling et al., 1980). In addition, it has been demonstrated that Ela can cooperate with the ras cellular oncogene to induce complete neoplastic transformation of primary rat embryo fibroblasts in culture (Ruley, 1983). 41 Both the Ela and Elb proteins form complexes with specific cellular proteins which have been shown to be involved in oncogenic transformation. Ela, like the large T antigen of SV40 and the E7 protein of HPV-16, forms a complex with the RB protein (Whyte et al., 1988). The EIB protein forms a complex with the p53 suppressor protein (Sarnow et al.,1982). It is believed that the formation of these protein complexes is an important function in the oncogenic activity of these proteins. c. Oncogenic Herpesviruses Epidemiological data have demonstrated a strong association with the infection of Epstein-Barr virus (EBV) and the development of several types of human lymphomas (e.g. Burkitt's lymphoma) and human nasopharyngeal carcinoma (Cotran et al., 1989). Burkitt's lymphoma is endemic in certain parts of Africa and New Guinea, whereas nasopharyngeal carcinoma is endemic to parts of Africa, Southern China and in Arctic Eskimos. Nearly 100% of patients in those areas of the world with these diseases carry parts of the EBV genome in their cells and have elevated antibody titers to the capsid proteins. Recent work in several laboratories suggests that the Epstein-Barr virus carries transforming genes. Elliott Kieff's group has shown that the viral latent membrane protein (LMP) gene has transforming activity (cited by Man, 1989). The mechanism of action for this gene is presently unknown, but it is thought to code for a growth factor or ion channel receptor. Another gene, EBNA-Z (Epstein-Barr virus nuclear antigen-2), was shown by William 42 Sugden to 'induce immortalization of lymphocytes (cited by Marx, 1989). This may be accomplished in part by stimulating the expression of other viral and cellular genes including the LMP gene. Additional genes or factors must be involved in the development of these cancers besides EBV infection. Most adults throughout the world are infected by EBV but outside of these particular geographic areas they are asymptomatic for these cancers (Cotran et al., 1989). One such factor in the case of Burkitt's lymphoma may be the requirement for specific genetic rearrangements. All of the patients with this disease have a chromosomal translocation of the distal part of chromosome 8 with either chromosomes 14, 22, or 2 (Klein, 1983). As will be discussed later, these translocations activate the expression of the cellular oncogene c-myc. d. DNA tumor viruses without oncogenes There are several oncogenic viruses which do not have any known transforming genes. One of these, the hepatitis B virus (HBV), has been closely associated with the occurrence of human liver cancer (Cotran et al., 1989). Like the Epstein Barr virus, HBV is endemic in many third world countries particularly those in Africa and in the Far East. Correspondingly these areas have very high incidences of hepatocellular carcinomas. A 10 year epidemiological study conducted in Taiwan has shown that carriers of the virus have about a 100 times higher probability of developing liver cancer than do uninfected individuals (United States Public Health Service Document, 1989). This is a significantly high risk when it is compared to 43 other cancer risk factors. For example, the epidemiological data are very conclusive in demonstrating that cigarette smoking is an important etiologic factor contributing to the development of lung cancer (Doll and Hill, 1950). Even as strong as this association is, the risk of developing lung cancer in smokers is still 5 times lower than the risk of HBV infected individuals to develop liver cancer. This level of risk leaves very little doubt that HBV is a major cause of this disease. Although transforming genes have not been detected in this virus several mechanisms have been proposed to explain its' oncogenicity. Work by Pierre Tiollais has found that sometimes when the virus inserts itself into the host genome it does so near a cellular oncogene. As a result this gene is inadvertently activated leading to its' unscheduled expression (Marx, 1989). Chronic liver damage, as a result of persistent HBV infection, also appears to play a important role in the tumorigenicity of this virus (Chisari et al., 1989). Transgenic mice were developed carrying the gene for the viral surface antigen. Expression of this gene in these animals resulted in large deposits of the antigen protein within the liver causing necrosis. This damage was accompanied by an increase in cell division to regenerate the damaged tissue. After months of tissue damage and regeneration, cancerous nodules began to develop in the liver of these animals. Although a build-up of this antigen does not occur in humans, chronic infection does cause liver injury as a result of immune attack on infected cells. The mechanism by which this virally induced repeated cellular division causes liver tumor development may be very similar to that discused above with certain nongenotoxic chemical carcinogens. '3.ClllQ 'C' . The identification of transforming genes in animal tumor viruses and the detection of sequences homologous to the retroviral oncogenes in normal mammalian cells suggested that neoplastically transformed cells may contain activated oncogenes. Experiments conducted in the late 1970's by Dr. Robert Weinberg's laboratory provided the first evidence for the existence of these cellular transforming genes (Shih et al., 1979). DNA was isolated from several chemically transformed rodent cell lines and when transfected into NIH 3T3 cells caused the morphological transformation of these cells. These experiments established the presence of oncogenic sequences within the DNA of the chemically transformed donor cell lines. Since these pioneering experiments, numerous studies have demonstrated the presence of transforming sequences in a variety of animal and human tumors (Bishop, 1987). Many of these transforming sequences have been identified to be homologous to the retroviral transforming genes. Other transforming genes found within these tumors have no apparent viral counterpart and therefore represent unique cellular oncogenes (Table 1). DNA sequences homologous to the cellular oncogenes have been found in normal mammalian cells and in organisms as evolutionarily 45 divergent as drosophila and yeast (Shilo and Weinberg, 1981; Powers et al., 1984; Chang et al., 1982). This suggests that the transforming oncogenes are derived from normal cellular genes (proto-oncogenes). The high degree of evolutionary conservation also indicates that the proto—oncogenes play a basic role in the cells normal biochemistry. These concepts raise several very irriportant and fundamental questions concerning the role and function of cellular oncogenes in the development of cancer: 1) how are the proto- oncogenes converted into their oncogenic counterparts and 2) what is the normal function of the proto-oncogene protein products and how is this altered by conversion to a transforming oncogene? Three basic mechanisms have been identified describing how proto-oncogenes are converted into transforming oncogenes: (1) point mutations (2) gene rearrangements (3) gene amplification. a. Activation by mint mutation The ras oncogenes represent the first and most extensively studied example of a family of proto-oncogenes activated by a point mutation. Point mutations activating this family of oncogenes (H, K and N) have been detected in three distinct functional regions within the gene. Mutations affecting codons 12 and 13, 61 and 117 have been detected in various human tumors and tumor cell lines as well as in experimentally induced animal tumors (Barbacid, 1987; Reynolds et al., 1987). Experimentally induced mutations in codons 59, 63, 116 and 119 have also been demonstrated to cause the activation of the ras gene (Fosano et al., 1984; Walter et al., 1986; Sigel et al., 1986). A single amino acid change at one of these codons is sufficient to convert the physiologically normal 21,000 dalton ras protein into its oncogenic form. The neu oncogene is a recent example of another oncogene which is activated by a point mutation. This gene was originally isolated from transformed NIH 3T3 cells that were transfected with DNA from cell lines derived from ethylnitrosourea induced rat neuro/glioblastomas (Schechter et al., 1984). The neu oncogene is homologous to, but distinct from, the gene coding for the epidermal growth factor (EGF) receptor. DNA sequence analysis of four independently derived NIH 3T3 transformed clones has revealed a single point mutation occuring at the same codon position within this gene (Bargman et al., 1986). This mutation results in the substitution of the normal valine with a glutamic acid within the transmembrane domain of the protein. Unlike the ras gene where an activating point mutation can occur at several different regions, the activation of the neu proto-oncogene apparently involves just a single site. The mechanism by which this mutation activates the neu gene is not understood at this time. Significant progress has, however, been made in understanding the transforming activity of the ras gene. This will be described later in more detail. b. Activation by gene rearrangement Specific chromosomal translocations occur at high frequencies in certain types of tumors, suggesting that these chromosomal rearrangements 47 may have an etiologic role in the development of these tumors (Brodeur, 1986a). This concept has gained greater acceptance with the finding that in several of these cancers the translocations involve the rearrangement of specific oncogenes. These rearrangements can result in the abnormal expression of the translocated gene or lead to the genesis of a hybrid gene whose protein product has transforming activity. One of the best examples involving this type of oncogene activation involves Burkitt's lymphoma. One hundred percent of the patients with this type of cancer have one of three translocations involving chromosome 8 (Cotran et al., 1989). The most common rearrangement involves a reciprocal translocation between chromosome 8 and chromosome 14. The c-myc oncogene has been located on the distal end of chromosome 8 where in normal cells expression is regulated during the cell cycle (Campisi et al., 1984; Dalla-Favera et al., 1982). This translocation results in the placement of the c- myc oncogene near the heavy chain irnmunoglobin gene (Taub et al., 1982). At this new chromosomal location the expression of the c-myc gene is under the control of the heavy chain gene. As a result high levels of c-myc expression are detected in these cells (Erickson et al., 1983a). Two alternative translocations involving chromosome 8 are observed in variants of Burkitt's lymphoma. In these variant translocations the c-myc gene remains on chromosome 8 and the immunoglobin kappa light chain gene on chromosome 2 or the lambda light chain gene on chromosome 22 are translocated and positioned near the c-myc gene (Erikson et al., 1983b; Hollis et al., 1984). In both cases the c-myc gene comes under control of the irnmunoglobin genes where transcription is greatly increased. A similar translocation is observed in acute T-cell or chronic lyniphocytic leukemia. Deregulation of the c-myc gene occurs when the T cell receptor on chromosome 14 is translocated near the c-myc gene on chromosome 8 (Erikson et al., 1986). A somewhat different type of oncogene activation involving a chromosomal rearrangement is found in chronic myelogenous leukemia. Greater than 95% of these cancers involve a reciprocal translocation between chromosome 9 and chromosome 22 (Nishimura and Sekiya, 1987). This translocation results in an abbreviated form of chromosome 22 which can be used as a prominent cytological marker (referred to as the Philadelphia chromosome) for the clinical diagnosis of the disease. The c-abl gene has been located on chromosome 9. This particular translocation joins the c-abl gene with the bar (breakpoint cluster region) gene on chromosome 22 to form a new hybrid gene (Heisterkamp et al., 1983). The normal c-abl gene shares sequence homology with the tyrosine kinase family of oncogenes but the protein product of this gene does not have any kinase activity (Ponticelli et al., 1982). The c-abI-bcr hybrid produces a chimeric protein with kinase activity. As a result of this specific translocation a new gene has been generated whose protein product has the biochemical activity of several other known oncogenes. c. Activation through gene amplification 49 The above discussion provided several examples of how increased oncogene expression, as a result of chromosomal translocations, could be associated with neoplastic development. Gene amplification is another mechanism in which high levels of oncogene transcriptional activity can be achieVed. This form of oncogene activation appears to be an important step in the genesis of certain types of malignancies. ‘ - For example, in approximately 40% of human neuroblastomas, the c- myc gene is amplified between 3 and 300 times (Brodeur, 1986b). There is a strong correlation between the degree of neoplastic progression and clinical prognosis with the extent of c-myc expression. The amplification of the c-myc gene has also been observed in neuroendocrine cells isolated from human colon carcinomas (Alitalo et al., 1983). Amplification of the neu oncogene has been detected in approximately 30% of human breast cancers (Slamon et al., 1987). As was observed with the c-myc gene, overall survival as well as the probability of remission is correlated to the degree of amplification. One possible explanation as to how amplification of this gene is related to these clinical parameters is provide by studies which have demonstrated that breast carcinoma cells expressing the epidermal growth factor receptor have the same poor prognosis (Sainsbury et al., 1987). Presumably the same ligand that binds to the EGF receptor also binds to the product of the neu oncogene (which is similar to the EGF receptor). Many breast carcinoma cells also produce transforming growth factor a (TGFva) which can bind to the EGF and neu receptor to stimulate growth. Synthesis of TGF-a by the breast carcinoma cells and its binding to 50 the neu receptor protein may set up an autocrine growth stimulatory process that contributes to the negative clinical prognosis associated with the amplification of the neu gene. Amplification of these genes can occur within the confines of the chromosome or externally to it (Schwab et al., 1983). The internal sites of amplification are often located in areas represented as homogeneously staining regions (HSR's). External amplification is found in extra- chromosomal fragments referred to as double minute chromosomes. Both types of amplification can visibly be observed through the cytological examination of cells isolated from the tumors. 4. W a. The retinoblastoma paradigm The above discussion has focused on a group of dominant acting oncogenes that cause neoplastic transformation, in part, by promoting excessive or inappropriate cell proliferation. These genes are considered dominant because they can induce neoplastic transformation when only one of the two normal proto-oncogene alleles is converted to its oncogenic form. If neoplastic transformation can be induced by the activation of a set of growth-promoting genes then, alternatively, cancers could arise by the inactivation of genes that normally function to suppress cell proliferation. Evidence to support this idea has been provided by a number of cell fusion experiments (Klinger, 1982; Klein, 1987). In these experiments cancer 51 cells are fused with normal cells and frequently the tumorigenic phenotype is lost in the hybrid cell. These results suggest that at least one of the properties associated with the neoplastic transformation of a cell involves the loss of growth-regulatory genes which are restored by the fusion to a normal cell. It is inferred from these experiments that normal cells contain specific genes that function to suppress properties associated with neoplastic growth. A single normal allele from one of these genes is all that would be necessary to exert this suppressive activity (unless there is a dosage effect). Consequently, both alleles must be inactivated if it is to participate in the neoplastic conversion of the cell. Because of this, these genes are sometimes referred to as recessive cancer genes (Hansen and Cavenee, 1988). Convincing evidence supporting the existence of cancer suppressor genes has been provided by the elegant studies determining the mechanisms involved in the causation of retinoblastoma. This particular type of cancer is relatively rare and affects approximately 1 out of every 20,000 infants and children (Cotran et al., 1989). The disease occurs sporadically in about 60% of the cases and in the remaining 40% there is a hereditary predisposition toward the development of the disease. Nearly 20 years ago, Knudson hypothesized that the development of this disease required two genetic events (Knudson, 1971). He suggested that in the hereditary cases, the first event is acquired from one of the parents. Careful cytological analysis of cells from individuals with the hereditary form of the disease occasionally revealed a small deletion in one of the number 13 chromosomes (Yunis and Ramsay, 1979). It was hypothesized that a gene 52 (referred to as the Rb gene) was located in the area of this deletion and that the inactivation of the Rb gene was involved in the development of retinoblastoma. The inactivation of the single Rb allele is however not sufficient to cause the disease. Only focal lesions are observed within the retina of individuals with the hereditary form even though all of the cells contain this initial defect. This suggests, as Knudson had hypothesized, that a second genetic event is necessary. This observation also supported his postulate that one of these events is genetically transmitted in the familial form of the disease and that in the sporadic cases two somatic events must occur within the same cell. Evidence supporting the existence of the second event was obtained by detecting DNA restriction fragment length polymorphisms between normal and tumor tissue at the Rb locus of individuals with the hereditary form of the disease (Sparkes et al., 1983). This analysis revealed that in the normal tissue of the retinoblastoma patient the Rb locus was heterozygous but in the tumor cells this locus existed in the homozygous condition. This established the fact that not only were two genetic events required but that these events involved the inactivation of both alleles from the same gene. Because both alleles of the Rb gene need to be inactivated to cause retinoblastoma, it was concluded that the Rb gene normally functions to suppress the tumor phenotype. These studies provided convincing evidence for the existence of tumor suppressor genes. b. Other possible tumor suppressor genes 53 A specific chromosomal deletion is also associated with Wilm's tumors. In this disease, the deletion occurs in the p13 band of chromosome 11 (Cotran et al., 1989). An experimental approach similar to the one described in the retinoblastoma studies has indicated that loci homozygocity also exists at the p13 region of chromosome 11 in Wilm's tumor cells (van Heyningen et al., 1985). Although the specific gene responsible for this disease has not been isolated, a recessive suppressor gene is suspected because of the similarity in the observations to the retinoblastoma example. A recent report by Levine and his colleagues provides data implicating the p53 gene as a possible suppressor gene (Findlay et al., 1989). In this study they generated transformed primary rat embryo fibroblasts by cotransfecting cultures with various combinations of oncogenes. For example, transformed foci were generated by transfection with the adenovirus EM and the activated ras oncogenes. If these primary cells were transfected with EIa, ras and with the wild type p53 gene, a significant reduction (67%-80%) in the number of transformed foci was observed. This result suggests that the p53 gene suppressed the transforming potential of the E1 a and ms genes. The incomplete suppression of foci development by the p53 gene was explained in the subsequent analysis of the few transformed foci that did develop. This analysis revealed that the p53 gene had been integrated into the genome of these transfectants but that it either was either not being expressed or was expressed in a mutant form. Because the p53 gene was basically inactive it could not exert its suppressive activity in those cells. Additional evidence supporting the idea that the p53 gene is a - suppressor gene is provided by the studies examining the interaction of this gene with the products several DNA tumor viral oncogenes. As was discussed earlier, these viral transforming proteins form stable complexes with the p53 protein. It is thought that the transforming activity of these viral proteins is due in part to the inactivation of the p53 protein through the formation of this complex. Such a mechanism would provide a very plausible explanation for the transforming activity of these viruses if the p53 gene has a suppressor function. This data, in conjunction with the information presented above, strongly suggests that the p53 gene may act as a suppressor of neoplastic transformation. ummmmmwm The regulation of normal cellular proliferation involves a complex interaction between extracellular and intracellular proteins. Information present in specific extracellular proteins is transmitted from outside the cell through a series of biochemical reactions into the nucleus which in turn instructs the cell to divide. Many of the viral and cellular oncogenes appear to play an important role in this information transfer. The protein products of these genes have altered biochemical activity which can interfere with the regulated transfer of this information resulting in the abnormal proliferation commonly associated with the ne0plastic cell. These oncogenic proteins can 55 be grouped according to their subcellular localization and biochemical activity (Table 1). 1. Wins An important step linking the protein products of oncogenes with aberrations in growth regulation was the finding that the simian sarcoma virus oncogene, v-sis, coded for a protein that was similar to the [3 chain of human platelet-derived growth factor (PDGF) (Waterfield et al., 1983). PDGF is a potent serum rnitogen that can induce DNA synthesis and cell division in many cells in the absence of other growth factors (Goustin et al., 1986). Several studies have shown that most transformed mesenchymal cells produce PDGF (Bowen-Pope et al., 1984). The biologically active form of PDGF consists of a dimer composed of a 14-18 kDa a chain and a 16 kDa [3 chain (Iohnsson et al., 1982). Stimulation of cells with PDGF results in the autophosphorylation of the 185 kDa PDGF receptor protein (Ek and Heldin, 1982). This presumably initiates a cascade of chemical reactions transmitting the proliferative information of PDGF to the cell nucleus. The transforming activity of the simian sarcoma virus is thought to be caused by the high level of. v-sis gene expression. This high expression of the v-sis gene results in the overproduction of PDGF which is believed to stimulate the tumor cell to grow by interaction with its own receptor through an autocrine mechanism (Sporn and Todaro, 1980). Evidence contradicting this autocrine hypothesis has been provided by studies where high concentrations of anti-PDGF antibodies added to the culture medium did not reverse the transformed phenotype of SSV-transformed cells (Huang et al., 1984). Furthermore, continuous addition of PDGF to the cultures of normal NRK cells does not induce cellular transformation. This contradiction may in part be explained by recent evidence where a mutant form of the v-sis gene was constructed which causes the v-sis protein product to be retained within the cell (Bejcek et al., 1989). PDGF produced by this gene is sequestered within the endoplasmic reticulum and the Golgi. Cells expressing this gene were transformed to the same extent as were cells transformed with the wild type v-sis gene. This data supports a mechanism whereby the v-sis and possibly the c-sis gene products may internally stimulate the PDGF receptor. 2. WWW There are several oncogenes, both of viral and of cellular origin, that code for growth factor receptor proteins located within the cytoplasmic membrane. These receptor proteins can be divided into three functional domains; an external domain which binds the growth factor protein, a transmembrane domain and a cytoplasmic domain having protein kinase activity. The first oncogene identified to code for a growth factor receptor was the transforming gene of the avian erythroblastosis virus (AEV), v-erb-B. Using a computer protein sequence data base, Downward et al. (1984) found that the the v-erb-B protein has extensive homology to the human epidermal 57 growth factor (EGF) receptor. The protein synthesized from the v-erb -B gene was however a truncated form of its cellular counterpart. This protein lacked a complete external domain which binds the EGF protein. The authors hypothesized that the transforming activity of the v- erb -B gene may be caused by a continuous proliferative signal elicited by the truncated receptor even in the absence of EGF. This hypothesis is partially substantiated by studies showing that AEV transformed erythroblasts fail to bind EGF (W aterfield, 1986). The myloid cell growth factor, CSF-l, is another example of a growth factor that is coded for by a potential cellular oncogene. This protein is synthesized by cellular proto-oncogene fms (Sher et al., 1985). A truncated version of this receptor protein is coded by the feline sarcoma virus oncogene v-fms. It seems that during the viral transduction of the cellular fms gene part of the 3' end of the gene was deleted (Waterfield, 1986). As a result the carboxyl terminal end of the v-fms protein is missing. This region of the protein represents the internal cytoplasmic domain of the molecule. The protein kinase activity of the protein is located in this portion of the receptor. It is thought that the transforming activity of this oncogene product is related to a possible loss of kinase activity. The neu, cellular oncogene codes for a protein which may represent a new growth factor receptor protein. The amino acid sequence of this protein is similar to but distinct from the EGF receptor. Unlike the erb-B gene, the neu oncogene has no viral counterpart. The activation of this gene also differs from that of the v-erb-B gene. Activation involves a single point 58 mutation in the transmembrane domain of the protein (Bargman et al., 1986). How this single mutation activates this gene is presently unknown. Many human tumors, particularly breast tumors, contain amplified copies of this proto—oncogene. In addition to the point mutation, an increase in gene expression as a result of amplification of this gene may play a role in its tumorigenicity (Slamon et al., 1987). 3. W The majority of the viral and cellular oncogene proteins reside within the cytoplasm however many are closely associated with the cytoplasmic membrane (Table 1). In general, the biochemical activity and role that these proteins play in cellular transformation are poorly understood. Many of these proteins have phosphokinase activity (Teich, 1986; Cotran, 1989). The majority of these proteins exclusively phosphorylate tyrosine although there are several that can phosphorylate serine and threonine (ie. mos, raf). These proteins also have the capability of phosphorylating themselves. This process of autophosphorylation may serve an important regulatory function in controlling their biochemical activity. One of the best studied oncogenic proteins within this group is the src protein. The gene coding for the src protein was originally identified as the transforming gene of the Rous avian sarcoma virus (Duesberg and Vogt, 1970). A homologous sequence, c-src, was subsequently identified and isolated in mammalian cells (Bishop, 1983). Activation of the c-src gene during viral transduction involves both genetic substitutions and point 59 mutations (Takeya and Hanafusa, 1983). The v-src gene codes for a protein with a molecular weight of 60,000 daltons and is usually referred to as pp605'c. It is believed that the transforming ability of this protein involves aberrations in the specific tyrosine kinase activity (Piwnica-Worms et al., 1987). A number of cellular proteins with phosphorylated tyrosine residues have been identified in cells transformed with v-src. One that is of particular interest is vinculin, a cytoskeletal protein (Cotran et al., 1989). This protein is a component of the cytoskeletal network of mesenchymal cells and is located at the points of contact between cells known as focal adhesion plaques. Changes in the phoshorylation of vinculin may be associated with some of the morphological changes detected in transformed cells such as rounding up and decreased cohesiveness. Some evidence also exists suggesting that pp605l‘c may also be capable of phosphorylating certain lipids (Cotran et al., 1989). Increased cellular proliferation by an activated src protein may involve a stimulation of the phosphotidylinositol signal transduction pathway. Another important group of cytoplasmic proteins, which lack phosphokinase activity, are the proteins coded for by the ras family of oncogenes. In light of the particular relevance of these proteins to this thesis, this class of oncogenic proteins will be examined in a separate section following the present general discussion of oncogene proteins. 60 4.:2 I'lll’ll'll 1 Products of several viral and cellular oncogenes are located within the nucleus. These genes include myc, fos, jun, myb, ets and ski (Table 1). Very little information is known about the ets and ski oncogenes. On the other hand the myc, fos, fun, and myb oncogenes have been studied more extensively. In general, the protein products of these oncogenes bind to DNA. Because of this, it is believed that they are involved in the regulation of gene expression either by activating or inactivating a particular gene or set of genes that participate in cellular proliferation or differentiation. Aberrations in this gene regulatory activity are thought to contribute to the transforming potential of these genes. This is exemplified by the c-myc gene protein. In humans this protein has 439 amino acids and can bind to single or double stranded DNA (Persson and Leder, 1984; Stewart et al., 1984). It is generally expressed at low levels in quiescent cells but expression is significantly increased after stimulation by certain growth factors or mitogens (Kelly et al., 1983). Regulated expression is absent in several chemically transformed cells where c—myc transcription is observed (Campisi et al., 1984). A viral homolog to the c-myc gene has been isolated from 4 different avian retroviruses that cause various sarcomas, carcinomas and leukemias (Klein, 1982). As discussed previously, chromosomal translocations involving this gene have also been associated with the causation of human Burkitt's lymphoma. In addition to the cellular myc gene, several related genes, N-myc 61 and L-myc have been isolated from human neuroblastomas and small cell lung carcinomas respectively (Klein, 1982). Interestingly, the v-myc or the c-myc gene when linked to a transcriptional promoter cannot neoplastically transform cells in culture. Instead transfection with these genes leads to cellular immortality (Mougneau et al., 1983). Transfection of quiescent cells with c-myc does not cause proliferation but instead prepares the cell for other growth signals. Consequently the myc gene is referred to as a competence gene which requires the action of other progression genes to complete the transformation process. This concept is illustrated in experiments conducted by Land et al. (1983) who showed that both the activated myc and ras genes were required to transform primary rat embryo fibroblasts in culture. Although considerable efforts have been extended toward understanding the role the myc genes play in cellular transformation, other than its DNA binding properties, the biochemical function of the myc protein still remains obscure. An interesting relationship between the protein products of the fos and jun oncogenes has been revealed during the past several years. The fos and jun genes were originally identified as the transforming genes of the FBI murine sarcoma (FBI-MuSV) virus and the avian sarcoma virus-17 respectively (Curran and Teich, 1982; Maki et al., 1987). Cellular homologues to both viral genes have also been identified (Van Beveren et al., 1983; Bohmann et al., 1987). Both v—fos and c-fos genes code for a 55,000 dalton protein which is localized in the nucleus (Curran et al., 1984). The pSSC' 5 protein differs from 62 the p55V'f05 protein by 49 amino acids at the carboxy terminus and it is also extensively post-transcriptionally modified. The fos protein, by itself, does not appear to have any DNA binding activity in vitro. However, antibodies to c-fos precipitate a protein complex which can bind to specific gene transcriptional control elements (Distal et al., 1987). One protein identified within this complex is a 39,000 dalton (p39) protein which has been identified to be the protein product of the c-jun oncogene (SassoneCorsi et al., 1988) Amino acid sequence analysis of the c—jun protein has revealed significant homology with the yeast transcription factor GCN4 (Vogt et al., 1987). This sequence homology reflects a functional similarity as both proteins bind to the same DNA sequence, TGACTCA (Hope and Struhl, 1985; Hill et al., 1986). The mammalian transcription factor AP-l also recognizes this sequence suggesting that the jun protein may be associated with AP-l (Angel et al., 1987; Lee et al., 1987). Using antibody cross reactivity reactions together with partial peptide sequence analysis it has been established that the c-jun protein is a major polypeptide of the AP-1 transcription complex (Rauscher et al., 1988; Sassone-Corsi et al., 1988). Separate in vitro translation studies with c-jun and c-fos have demonstrated that the c-jun protein forms a homodimer which binds to the AP-l DNA site while the c-fos protein fails to dimerize and does not bind to DNA (Halazonetis et al., 1988). However if both genes are cotranslated, a heterodirner is formed that binds approximately 25 times more efficiently to the AP-l site" than does the c-jun homodimer. These experiments suggest that the transcriptional activation of genes by AP-1 requires the coordinated 63 synthesis and interaction of the c—jun and c-fos proteins. If there was an over expression of the c-jun gene in the absence of c-fos expression then a predominant formation of a c-jun homodimer may occur. Because this homodimer binds less efficiently to the AP-l binding region compared to the jun lfos heterodirner it could cause the deregulation in the control of gene transcription. This may be one mechanism explaining how these genes are involved in neoplastic transformation. BMW 1. Weaning: The mammalian ras genes are members of a multigene family (H-ras, K-ras, and N-ras) which have been implicated in playing an etiologic role in most human tumors as well as in a variety of spontaneous and experimentally induced animal tumors (Barbacid, 1987; B05, 1988). Two ras genes, H and K, are homologous to the retroviral transforming genes of the Harvey and Kirstin murine sarcoma viruses. The N-ras gene was isolated from a human neuroblastoma cell line and does not have a viral counterpart (Shimizu et al., 1983). Sequences homologous to the mammalian ras genes have been detected in yeast and drosophila (Shilo and Weinberg, 1981; Chang et al., 1982). The conservation of these genes throughout eucaryotic evolution suggests that they play an intrinsic role in basic cellular physiology. All 3 ras genes contain a 5' non-coding exon followed by 4 coding exons which code for a very similar 21 kDalton protein (Pulciani et al., 1982; McGrath et al., 1983; Cichutek and Duesberg, 1986). The K—ras gene can produce several variations of the ras protein which differ primarily at the carboxyl terminus. These variations arise as a result of alternative mRN A spicing patterns transcribed from two different number four exons (McGrath et al., 1983; Capon et al., 1983, Fasano et al., 1983). The ras proto—oncogenes can be activated by the induction of specific point mutations within the coding exons of the gene. All of these mutations are clustered within specific codons in exons 1, 2 and 3. Mutations within codons 12, 13, 61 and 117 have been detected in naturally occuring human and in chemically induced animal tumors (B05, 1989; Reynolds et al., 1987). Earlier studies have indicated that only 10-20% of human tumors contained activated ras genes (Pulciani et al., 1982). With the development of more sensitive and rapid techniques of analysis, this frequency has increased substantially for some types of tumors. For example, more than 90% of human pancreatic tumors have been shown to contain an activated K-ras gene (Smit et al.,1988). In vitro mutagenesis of the ras genes has revealed several other activating codon positions. Treatment of the H-ras gene with bisulfite induces mutations in codons 59 and 63 that can transform transfected NIH 3T3 cells in culture (Fasano et al., 1984). Site directed mutagenesis experiments have also indicated that changes in the coding sequence at codons 116 and 119 have similar transforming capabilities (Walter et al., 1986; Sigel et al., 1986). Most of these natural and synthetic mutations involve single base pair substitutions within the specific activating codon. A double base pair substitution has been detected in the 12th codon of the H-ras gene form Balb-C murine sarcoma virus (Reddy et al., 1985). The normal mouse nucleotide sequence at this position is a GGA coding for a glycine and in the virus this codon has been mutated to an AAA which codes for a lysine. Site directed mutagenesis experiments by Seeburg (1984) have indicated that mutations substituting any amino acid except proline at position 12 will activate the H-ras gene. Some mutations were observed to be more potent than others as determined by in vitro transfection analysis into rat-1 cells. Activation of the ras genes can also occur as a result of increased gene expression. Multiple copies of the H-ras gene transfected into NIH 3T3 mouse fibroblasts‘cause the morphological transformation of these cells (Pulciani et al., 1985). Fusion of strong viral promoters to the normal H-ras gene leads to the enhanced expression of this gene which can transform cells in culture (Chang et al., 1982). In addition, elevated expression of the human ras genes have been frequently detected in premalignant and malignant tumors of the colon and to a lesser extent in the lung (Spandidos and Kerr, 1984; Pulciani et al., 1985). There have been conflicting reports addressing the issue as to whether the ras genes alone are sufficient to cause neoplastic transformation. Studies using primary rodent fibroblasts in culture have demonstrated that at least one other oncogene is required for tumorigenic conversion by the ras genes. Cellular immortalization appears to be an important prerequisite for this transformation. Prior immortalization by exposure to mutagenic chemicals or transfection with either the myc or the adenovirus EIa gene was necessary for successful transformation of primary rodent cells by ras (Land et al., 1983; Newbold and Overell, 1983; Ruley, 1983). ' Multiple gene activation also appears to be a requirement for the complete transformation of diploid human fibroblasts by the ras oncogenes. Studies by Hurlin et al. (1987) and Wilson et al. (1987) have demonstrated that the transfection of normal human fibroblasts by the H-ras and N-ras genes resulted in these cells acquiring several characteristics of transformed cells (ie. focus formation and anchorage independence) but failed to exhibit an infinite lifespan in culture and were not tumorigenic. If however an immortalized human cell line was used in these experiments then complete neoplastic transformation was achieved by transfection with the H-ras gene (Hurlin et al., 1989; Wilson et al., 1989). In vivo evidence in support of the requirement for multiple gene activation for neoplastic transformation is provided by studies where transgenic mice have been created carrying the v-H-ras gene linked to the mouse mammary tumor virus promoter (Sinn et al., 1987). These mice developed tumors in the mammary, salivary and lymphoid tissues. The tumors that did develop in these tissues were however isolated focal lesions of monoclonal origin indicating that additional somatic events were required for full malignant transformation. Contrary to the results of these studies, Spandidos and Wilkie (1984) have reported the complete neoplastic transformation of early passage Chinese or Syrian hamster cells by the T24 H-ras oncogene linked to transcriptional enhancers. Transformation was dependent on both the 67 mutational activation and enhanced expression of the ras gene. Transfection of the H-ras gene without the enhancer elements resulted in immortalization only. These results were confirmed in a separate study using rat embryo cells transfected with the H-ras gene (Pozzati et al., 1986). One problem with the interpretation of these studies is that it has not been determined whether additional oncogenes were spontaneously activated in these cells. Consequently, it cannot definitively be established that the activation of a single oncogene is sufficient to induce complete neoplastic transformation. Indeed, the bulk of experimental evidence supports the concept that carcinogenesis is a multi-step process of which one may be the activation of the ras gene. 2. W The ras protein can be divided into four functional domains based on their biochemical activity and evolutionary conservation (see Figure 2). The first domain is represented by the first 80 amino acids where almost complete evolutionary conservation has occurred among the mammalian ras genes (Santos and Nebreda, 1989). Located at codons 12 and 61 within this domain are two areas where the majority of the activating mutations take place. These two areas also encompass two of the four GTP binding sites which are essential for oncogenic transformation by the protein (DeVos et al., 1988). Another essential region found within this domain occurs between amino acids 32 and 40. This region is not involved with GTP binding or hydrolysis hOwever any amino acid changes occurring at this site eliminate the transforming activity of the protein. 69 89 .3282 new 3.5% E0... potato Am 865m $8282 I 52.8 29885 _H_ 52.8 .285 a . 58.8 .9533 I NVFAVVF $2.-wa mmtmm . owtww . < < e < 4 4 4 4 1000 2: 8e 9; 8e 2: 8 8 9. ow . mEmEou .2285“. .n :99: 225283 a\e 8-2 o\o 3A :08 _ _ _ _ _ _ _ ~12 8. of 9; our 2: 8 8 9. 8 e cozmzmmcoo 9325.05 8 ocom ow: :EEEEmE 05 do mEmEou .mcozoca can 55625 .N 9.39.... 70 It has been suggested that this region may be involved in the interaction of the ras protein with its putative cellular target and consequently it has been named the effector domain (DeVos et al., 1988). The second domain encompasses amino acids 81-160 where approximately 80-90% amino acid sequence homology exists among the various mammalian ras proteins. Within this domain there are three essential regions ( a.a. 77-92, 109-123 and 139-165) that are absolutely necessary for cellular transformation. The other two GTP binding regions are located within two of these essential regions occuring between amino acids 116-119 and 143 -147. The remaining amino acids, excluding the last 4, constitute the third domain and are generally referred to as the hyper-varible portion of the protein. The amino acid sequence within this domain differs for each of the ras proteins. Most amino acid changes within this region do not appear to effect the transforming activity of the protein. The fourth domain is represented by the carboxy terminal four amino acids and is designated as the CAAX box where the C is a cysteine residue, the A can be any aliphatic amino acid and the X is any amino acid. This amino acid sequence is highly conserved in all of the ras proteins and is present on several other unrelated G proteins, nuclear lamins and the yeast alfa mating factor (Clarke et al., 1988). This motif plays an important role in the posttranslational modification of the ras proteins. 71 3.2 II I I' 1 IT I' [I] I . The newly synthesized ras proteins are located within the cytoplasm and are biochemically inactive. Functional activation requires the localization of the ras proteins at the plasma membrane. Transport to the membrane involves a series of complex posttranslational modifications. These modifications involve proteolytic cleavage, acylation, polyisoprenylation and palmitoylation of the carboxyl terminal portion of the protein. The terminal CAAX motif plays a pivotal role in this process. Early studies indicated that palmitoylation of the ras proteins was an important process for membrane localization (Chen et al., 1985; Willumsen et al., 1984). In these studies it was demonstrated that the replacement of the terminal Cys136 with serine resulted in the blocking of palmitoylation and membrane localization. These results suggested that the Cy5136 residue was the site of protein palmitoylation and that the CAAX box was the determinant for processing and localization of p21 ras. Very recent studies have shown however that Cys186 is not the site for palmitoylation and that the posttranslational processing of the ras proteins involves a two step process. Gutierrez et al. (1989) have demonstrated that during step 1 the initial translation product, pro-p21, is synthesized within the cytoplasm and rapidly converted to an intermediate form, c-p21. This conversion involves the proteolytic cleavage of the terminal AXX amino acids and the methylation of the a-carboxyl group on Cysl36. These modifications increase the hydrophobicity of the c-p21 compared to the pro- p21 but the protein is still located in the cytoplasm and is not palmitoylated. Subsequent studies by Hancock et al. (1989) have shown that the cytoplasmic ras proteins are also polyisoprenylated during this first step and that the site of this isoprenylation is the terminal Cys186 residue. During the second step the palmityl groups are added and the protein is localized at the membrane (Hancock et al., 1989). Palmitoylation of the cytoplasmic ras proteins occurs via thioesterification of palmitic acid with cysteine residues that are upstream of Cysl36. This takes place at Cys181 and Cys184 on H-ras, Cys181 on N—ras and presumably at Cys180 on the viral K-ras proteins. The p21 K(B)-ras mammalian protein is not palmitoylated because it lacks cysteine residues upstream of Cys‘86. Since the K-ras protein is not palmitoylated it must not be an essential modification for protein activity because this protein still has transforming activity. All ras proteins are therefore isoprenylated but not necessarily palmitoylated. Hancock et al. (1989) did show that palmitylation increases the avidity of membrane binding as well as the transforming activity of the ras proteins. The importance of polyisoprenylation for ras function is demonstrated in an experiment in which polyisoprenoid biosynthesis is inhibited by treatment of cells with mevinolin (Hancock et al., 1989). This blockage of polyisoprenylation abolished p21 membrane binding. Because membrane binding is essential for the biological activity of the ras proteins, polyisoprenylation is essential for cellular transformation. 73 4.111.] 'll"l [I] I' Extensive analysis of the protein products of the proto-oncogenic and oncogenic ras genes has yielded very little functional information in regards to their role in normal or neoplastic cellular processes. In spite of this lack of functional information, a considerable amount of data has been generated concerning the structure and biochemistry of these proteins. The ras proteins share many biochemical properties with a class of proteins, commonly referred to as the G-proteins, which are involved in membrane receptor mediated signal transduction pathways. These similarities suggest that the ras proteins may also function in this capacity and that aberrations in this process may explain their transforming activity. Like the G-proteins, the ras proteins have the ability to bind, exchange and hydrolyze guanine nucleotides (Scolnick et al., 1979; Hurley et al., 1984). Regulation of the biochemical activity of the G—proteins involves the cycling between the inactive GDP-bound form and the active GTP-bound form (Gilrnan, 1984). A similar mode of activation and inactivation is thought to occur with the ras proteins. It has been hypothesized that the transforming activity of the mutant ras proteins may be the result of abnormally high levels of p21 in the active GTP-bound state (Gibb et al., 1984). These mutations could induce high levels of GTP-bound p21 by either altering the GTP binding affinity of the molecule or by inhibiting the proteins intrinsic GTPase activity. Structural studies coupled with mutational analysis have determined that the binding sites for GTP are centered around the transformation activating ,codons 12 and 61 (DeVos et al., 1988). Mutations at these codon 74 positions however do not appear to manifest their transforming potential by effecting GTP binding because both the normal and transforming ras proteins bind GTP with similar affinities (Finkel et al., 1984). Comparative biochemical analysis between normal and the activated p21 protein has revealed significant decreases in the GTPase activity of various mutant ras proteins (Gibbs et al., 1984; McGrath et al., 1984; Manne et al., 1985). A correlation between the decrease in GTPase activity of these proteins with in vitra and in viva transforming potential however cannot be established (Lacal et al., 1986; Trahey et al., 1987a). The intrinsic GTPase activity of the normal p21 protein is very weak under normal physiologic conditions and is insufficient to maintain the ras protein in the inactive GDP state (Trahey and McCormick, 1987b). Recently a protein has been isolated which associates with p21 and stimulates its GTPase activity (Trahey and McCormick, 1987b). This GTPase activating protein (GAP) catalyzes the conversion of p21-GTP to p21-GDP at a rate which is several hundred fold greater than the intrinsic rate. It is of particular significance that this protein does not have a stimulatory effect on the GTPase activity of ras proteins with mutations in codons 12, 59 or 61 (Adari et al., 1988). Because many of these mutant proteins already have reduced GTPase activity, their in viva GTPase activity when coupled with GAP is several orders of magnitude lower than the wild type p21. Functional mutation studies have identified a region of the p21 protein located between amino acids 32 and 40 as a possible interaction site between the ras protein and its putative cellular target (DeVos et al., 1988). Mutations 75 within this region do not effect GTP binding, GTPase activity or membrane localization but is essential for ras transforming activity. This area has been termed the ras effector region. Recent studies have tested the ability of various mutant ras proteins to interact with GAP (Adari et al., 1988). Proteins with mutations in non- essential regions were stimulated by GAP but proteins with mutations in GTP binding regions or in the effector region lost their ability to respond to GAP. It was concluded that the ras effector region is the site for GAP binding and it has been suggested that GAP is the ras effector protein (McCormick, 1989). A model has been proposed by McCormick (1989) which postulates GAP as being both an effector and regulator of ras protein activity. In this model membrane bound p21-GDP is converted to p21-GTP as the result of some upstream stimulatory signal (for example it could be initiated by a growth factor). This nucleotide exchange results in the formation of a p21- GTP/ GAP complex. This complex may serve to position GAP at a specific site in the membrane where GAP can further interact with some other cellular factor. Thus, the input signal is propagated through a ras-GAP complex where GAP serves as the effector transmitting the stimulatory signal to the next component in the signal cascade. The duration of the signal is limited because p21-GTP is quickly converted to p21-GDP by GAP resulting in the disassociation of GAP from the ras protein. GAP therefore participates in the signal transmission (as an effector) and at the same time (as a regulator) down regulates the input signal . Transforming ras proteins still associate with GAP but the output signal is no longer down regulated because the ras GTPase is 76 not stimulated by GAP. GAP remains associated with ras and is properly positioned to participate in the signal cascade. Consequently, an unregulated, constitutive output signal is propagated in cells carrying mutant ras genes contributing to the transformed phenotype. EC] '1: . '3']! I 1. Methndsmedinriskassessmcnt The extensive use of synthetic chemicals by industrialized societies in conjunction with the possible carcinogenic hazard associated with exposure to these chemicals has created the need to predict and evaluate their carcinogenic potential. This process currently relies on data generated from a a number of experimental methodologies because no single test can equivocally determine a chemical's carcinogenic potential. This usually includes the use of human epidemiological studies in conjunction with various in vitra and in viva test systems. Data generated from all of these sources are assimilated to establish the carcinogenic potential of a particular chemical. It is important to recognize that even with the use of this multi- faceted approach the evaluation of a chemical's carcinogenic potential is still an approximation. This section will discuss the approach generally used to make this assessment and why the interpretation of results from these assays provides only an approximation rather than an absolute determination of carcinogenic risk. a.) Epidemiological analysis in risk assessment A fundamental objective of the cancer epidemiologist is to study the patterns of human cancers in an attempt to elucidate the various factors associated with its cause. This is basically achieved by describing the occurrence of cancers in relation to differences in a number of factors such as sex, age, race, geographical location, occupation and socio—economic class to determine a common parameter associated with the disease outcome. The utility of this approach has been validated by a number of studies correlating specific human cancers with exposure to a specific chemical agent. These studies have provided great insight into the various causes of cancer which has been beneficial in identifying and eliminating exposure to a number of hazardous materials. Perhaps one of the most successful examples of how epidemiology can contribute to the identification of a human carcinogen is provided by the studies which have identified cigarette smoking as an important etiologic agent in the development of lung cancer (Doll and Hill, 1950). An examination of cancer death statistics revealed an enormous increase in the lung cancer incidence in males, but not in females, between 1920 and 1944. This suggested that some exclusive activity of males was the cause of this increase. After careful analysis of a large volume of data the common factor of cigarette smoking was identified. One of the difficulties in identifying cigarette smoking as the etiologic factor for the increase in male lung cancers was the long latency period between when the individual began smoking and the onset of disease. This latent period averaged around 20 to 30 years from 78 the time of initial exposure to the appearance of a clinical tumor. The use of epidemiological analysis not only identified the causative agent for the increase in this particular cancer but it also provided an important insight into the mechanism of chemical carcinogenesis. Once the causative agent was determined, it was possible to initiate studies to identify the specific chemicals within tobacco smoke that were responsible for inducing the cancer. Many of these chemicals have been isolated and identified and include a number of polycyclic aromatic hydrocarbons such as benzo(a)pyrene (IARC Monograph, 1982). Using an approach similar to that used in the smoking and lung cancer studies, epidemiologists have identified a number of other chemicals as human carcinogens. This list includes such industrial chemicals as benzene, 4-aminobiphenyl, asbestos, benzidine, cyclophosPhamide and vinyl chloride in addition to several pharmaceutical agents such as diethylstilbestrol and other conjugated oestrogens (IARC Monograph, 1982). In spite of the utility of using epidemiological studies to evaluate human chemical carcinogenicity there are several important limitations to this approach (MacMahon, 1979; Muir, 1979). One of the most serious limitations is that epidemiology is primarily an observational science. Conclusions depend on collecting and comparing a large number of observations and then finding a common parameter associated with a particular disease state. It is often a very difficult task to accomplish this with a high degree of statistical and biological confidence. 79 These analyses are dependent on the availability of accurate personal records (ie medical, occupational, family histories etc.) to determine cause and effect relationships. Consequently inadequate records can severely hinder the interpretive conclusions by the investigator. Achieving a high degree of statistical confidence usually necessitates the analysis of fairly large population sizes. In many cases the exposed populations are very small and it is often very difficult to ascribe any statistically significant increase in tumor development with exposure. Also it often requires taking multiple measurements which is rarely possible. A greater degree of confidence is obtained if the disease is relatively- uncommon and the exposure causing the disease is unusual. This is illustrated by the epidemiological study associating the development of angiosarcoma of the liver with the exposure to vinyl chloride (Doll, 1988). Because the incidence of this disease is quite rare and the occurrence was confined to a specific occupational exposure group, a great deal of confidence was obtained in assigning the increase in this cancer with exposure to vinyl chloride. It is more difficult to correlate the increase in a disease with exposure when the the disease is fairly common and the natural incidence fluctuates with factors such as time, geographic location and socio-economic status (Muir, 1979). All of these factors limit the ability of the epidemiologist to evaluate the carcinogenic potential of any chemical agent. This is reflected in the fact that of the several hundred chemicals which have been assigned animal carcinogens less than 25 have been confirmed as human carcinogens by epidemiological analysis (IARC Monograph, 1982). Consequently our risk evaluation process has evolved to include an experimental approach to assist in the determination of human chemical carcinogenic potential. b.) Shag-m tgsts fgr thg gvaluatign 9f chemical sarg’ngggnig mtgntial The retrospective nature of epidemiological analysis severely limits the utility of these studies for predictive human carcinogenic risk evaluation. The animal bioassay was developed to provide a prospective experimental method to assess carcinogenic risk. These assays are, however, very time consuming and expensive to conduct. This has necessitated the development of alternative methods to screen for the thousands of chemicals with potential carcinogenic activity. To assist in this task a number of short-term tests for mutagenicity have been developed. The rationale for the use of a mutagenicity test as a predictor of carcinogenicity is based on the preponderance of data implicating mutagenicity with carcinogenic activity. In addition, the speed, relatively low cost and high predictive value has added to the appeal of these test systems and as a result they have become an important component in carcinogenicity evaluation. Because of specific interpretive limitations with any one assay, the evaluation process has evolved into the use of a battery of test systems. These testing protocols ideally would include assays to measure mutagenicity in 81 bacteria and mammalian cells in vitra and gross chromosomal damage both in. vitra and in viva (Wiliams et al., 1980). Of the over 100 short-term tests that exist the Ames Test is the most well known and widely used. This test assesses chemical mutagenicity using several specially developed strains of Salmonella bacteria (Ames et al., 1973). Each strain has been constructed to detect a chemical's potential to induce either base pair substitutions or frameshift mutations. Because many chemicals require bioactivation to their active mutagenic form, the test is conducted with and without a rat liver preparation that is rich in biometabolizing enzymes (59 mix). The bacteria themselves are auxotrophic mutants which cannot synthesize histidine. The basis of the test is to determine whether a chemical can induce mutations causing a reversion of the mutant bacteria to a wild type phenotype that can then be selected for on minimal media plates. This test detects most of the known organic human chemical carcinogens with the exception of benzene and diethylstilbesterol (IARC, 1980). An extensive analysis has been conducted to evaluate the predictive value of Ames tests results with animal bioassay carcinogenicity data (McCann et al., 1975 ). Of 254 positive carcinogens tested there was a 90% correlation between mutagenicity and carcinogenicity. In general, the metal carcinogens, hormones and nongenotoxic carcinogens were not detected in the test system. In the evaluation of noncarcinogens, no mutagenic activity was detected in 46 of 46 tested chemicals. 82 Because mammalian cell DNA is more complex than bacterial DNA, in vitra mammalian mutagenicity assays are also conducted. The basic design of these test systems is in principle very similar to the Ames test. The objective of these tests is to determine the ability of the chemical to induce mutations in a specific genetic marker in mammalian cells in culture. These markers typically utilize a gene involved in an easily selected for biochemical metabolic pathway . Two of the more commonly used systems are designed to detect mutations in either the hypoxanthine guanine phosphoribosyl transferase (HGPRT) gene or the thymidine kinase (TK) gene (Williams et al., 1980). . Some chemicals damage genetic material at the level of the chromosome and consequently would not be detected as mutagens using the assays described above. Chromosomal damage can be classified in two general categories (IARC, 1980). One involves the whole chromosome and the other involves the individual chromatids. Whole chromosomal aberrations occur during the G1 phase of the cell cycle prior to DNA replication. This damage involves breakages that can lead to gaps, terminal and interstitial deletions, inversions, centric and acentric ring formation, and asymmetric interchanges that can produce reciprocal translocations. Chromatid aberrations involve damage to the single chromatid which can occur at any time during or after the replication of the chromosome. A wider variety of damage can be observed with this type of aberration compared to the whole chromosome. Both in viva and in vitra tests have 83 been developed to evaluate a chemical's potential to induce either type of damage. The in vitra systems typically use peripheral blood lymphocytes or fibroblasts for analysis. Some of the advantages of the in vitra systems are that they can be conducted with human cells and that the concentrations of test material can be much higher than would be possible with in viva studies. A major disadvantage is that the cells in culture may be unable to metabolize the chemical to its mutagenic form. This is circumvented by the addition of a metabolizing system similar to what is used in the Ames test. The use of in viva systems is very desirable because the whole animal is exposed to the test material which allows the influence of bioavailability and metabolizing functions on mutagenicity to be evaluated. This makes it possible to determine quantitative measurements of dose response which is very important for risk assessment analysis (Williams, 1980). In viva analysis also allows for the analysis of a wider variety of cell types which can include peripheral lymphocytes and cells of the bone marrow, liver, gonads etc. (IARC, 1980). Genetic endpoints for clastogenic in viva and in vitra test systems include the cytological chromosomal abnormalities discussed above, the presence of micronulcei, and the determination of sister chromatid exchanges. Endpoints restricted to the in viva systems would include dominant lethal mutations and heritable translocations within the germ cells (IARC, 1980). One of the major disadvantage of these last two endpoints is that they are time consuming and labor intensive and as a result studies to evaluate them are very expensive to conduct. 'An obvious major limitation for the evaluation of chemical carcinogenicity using the short-term tests described above is that these assays are determining mutagenic potential rather than the ability to cause cancer. Even though there is a strong relationship between mutagenic potential and carcinogenic potential, the correlation is not 100%. In spite of these limitations the results from short-term assays continue to play an important role in the evaluation of a chemicals carcinogenic potential. c.) Thg animal bigassay Data generated from epidemiological and mutagenicity studies are important contributions in the evaluation of a chemical's carcinogenic potential. The results from the rodent bioassay for carcinogenicity however continues to have a major influence in this decision making process. The general premise of the long-term bioassay is to expose animals to a specific chemical for the majority of their lifespan and determine whether this exposure results in the increased incidence of neoplastic lesions. These studies are a time-consuming and very costly endeavor and it is important that a degree of uniformity be achieved in their conduct so results from different laboratories will have comparable interpretations. To achieve this objective various national and international committees have convened to adopt protocol guidelines for the conduct of these studies (WHO, 1969, 1978; FDA, 1971). i.) Bssic condugt The rodent bioassay, as it is currently conducted by the National Toxicology Program under the auspices of the National Institute of Environmental Health Sciences, was adopted during the early 1970's and has been used to evaluate over 300 chemicals for potential carcinogenicity. These studies are presently concluded according to international guidelines and conform to the principles of Good Laboratory Practices (Federal Register, 1978). These guidelines specify that males and females from two rodent species will be exposed to the test material for a 24 month period. The route of administration of the test material will be chosen to reflect the probable route of human exposure. Experimental groups will consist of a control group receiving no test material, a high dose group in which the exposure concentration is set at the maximum tolerated dose (MTD) and a low dose group where the exposure concentration is 1 / 2 the MTD. Each group is composed of 50 animals and the species primarily used are the B6C3F1 mouse and the F344 rat. At the termination of the study a chemical will be considered positive for carcinogenicity if a statistically significant increase is observed in the overall tumor incidence or in a specific tumor type in either sex in at least one of the two test species. ii) Problems with interpretation The development of animal tumors in the bioassay does not definitively establish a chemical as a human carcinogen. A number of factors complicate the interpretation and extrapolation of bioassay results. The specific influence that many of these factors have on the eventual outcome of bioassay results is not completely understood at this time. This makes the evaluation of human carcinogenic risk a difficult and uncertain task. I For example, in many bioassays there is a difference in the susceptibility to develop tumors between the 2 rodent species used. This is illustrated in an examination of 226 chemicals that were positive for tumor induction in the bioassay. Of these 226 bioassays, 42% (96 of 226) produced tumors in the mouse and not in the rat or vice versa (Bernstein et al., 1985). It is difficult to assign a carcinogenic risk in humans when such a disparity exists in the tumor response of two so closely related species. Metabolic and pharmacokinetic differences between rodents and higher mammals can also complicate the extrapolation of bioassay data. This is particularly important for the many compounds that must be biotransformed to a reactive carcinogenic intermediate. Differences in uptake and elimination rates can result in the saturation of primary detoxifying enzymatic pathways in one species and not in another. This can lead to the activation of secondary routes of metabolism which can produce reactive intermediates. These secondary metabolic intermediates may then interact with important macromolecules, such as DNA, and initiate neoplastic development. This problem is illustrated in rodents which tend to use glutathione rather than epoxide hydrase for detoxification of xenobiotics. Man and primates utilize glutathione only as a reserve after epoxide hydrase activity has been saturated (Pacifici et al., 1981). Consequently at high exposure 87 concentrations intracellular glutathione is depleted early in the detoxification process making rodents potentially more vulnerable to the generation of various oxidants and free radicals. These differences in metabolic routes of detoxification can influence the sensitivity of a species to develop tumors after exposure to a particular chemical. Such species differences must be considered when extrapolating bioassay results to assess human carcinogenic risk. The experimental design of the bioassay is also a contributing factor effecting the interpretation of results. As indicated previously, the protocol for the rodent bioassay calls for the exposure of a relatively small number of animals to very high doses of the test material. The use of a small number of animals is based on logistic and economic reasons and the use of high doses is to maximize the probability of inducing a biologic effect. As was previously discussed, high exposure concentrations may alter the metabolic fate of the test chemical and produce a response that would not be observed at the much lower exposure levels to which humans typically would be exposed. Many of the inbred strains of rodents have high spontaneous tumor incidences. For example nearly 100% of F344 male rats will develop spontaneous testicular tumors within two years (WHO, 1980). The Sprague- Dawley rat and the A / I mouse have high spontaneous incidences of mammary and lung tumors respectively. One of the more controversial factors complicating the evaluation of bioassay results involves the relevance of an increase in a normally high spontaneous tumor incidence to the potential human risk. This problem is particularly relevant with the B5C3F1 mouse because of its extensive use in many bioassays. Males of this strain have a characteristically high spontaneous liver tumor incidence which averages around 30% in control p0pulations but this incidence can vary from 10%-50% (Chan, 1984). In manylbioassays involving the 86C3F1 mouse an increase in this Mar incidence is often observed following treatment with the test compound. This is illustrated in an examination of 85 positive bioassays conducted by the National Cancer Institute between 1977 and 1980 involving the B5C3F1 mouse and either the F344 or Osborne Mendal rat (Hamm, 1984). Of the 85 chemicals that were carcinogenic in at least one sex of one species, 59% (50/85) caused liver tumors in the B6C3F1 mouse whereas only 22% (19/ 85) produced liver tumors in the rat. In addition, 24% (20/ 85) of the chemicals caused only liver tumors in the mouse and no tumors of any type were observed in either rat strain. The significance of the increase in only mouse liver tumors to a potential human carcinogenic risk from exposure to these chemicals is presently a topic of considerable debate. Numerous task forces have been assembled in an attempt to resolve this issue with little success. Because of the preponderance of data associating mutations with neoplastic development there is less controversy interpreting the significance of B5C3F1 mouse liver tumors induced with genotoxic chemicals. Since these agents are also likely to produce mutations in humans there is greater confidence in classifying them as potential human carcinogens. However, an increase incidence in the B6C3F1 mouse liver tumors has also been observed 89 for a number of chemicals that appear to have little if any mutagenic activity. For example, the chlorinated hydrocarbons tetrachloroethylene, trichloroethylene, tetrachloroethane and chloroform are generally negative in various short-term tests for genotoxicity yet all cause a significant increase in B6C3F1 mouse liver tumors (Environmental Health Criteria, 1985; Reichert, 1983). The debate concerning the significance of B6C3F1 mouse liver tumors to potential human carcinogenic risk is currently centered around bioassays involving compounds of this type. To accurately evaluate human carcinogenic risk from nongenotoxic chemicals it is important to understanding the mechanism by which these spontaneous tumors arise and how chemical agents enhance this tumor response. 90 GO ll'l' 'BCEM I' I 1.) MW The detection of activated cellular oncogenes in a variety of human tumors suggested that cellular oncogenes might also be associated with the induction of spontaneous liver tumors in the B5C3F1 mouse. The presence of dominant acting oncogenes can be detected using cell mediated DNA transfer techniques (Shih, 1979). To determine whether spontaneous 86C3F1 mouse liver tumors contained dominant cellular oncogenes several investigators transfected DNA isolated from these tumors into NIH 3T3 cells (Fox and Watanabe, 1985; Reynolds et al., 1986). These transfection experiments resulted in the development of morphologically transformed foci indicating the presence of a dominant oncogene(s) (Table 3). The percentage of animals whose tumor DNA induced transformed foci ranged from 56% to 82% suggesting that the activation of cellular oncogenes was an important event in the development of these spontaneous tumors. The B6C3F1 mouse is a hybrid strain resulting from the cross between a C3H/He] male and a C57BL/6I female. The C3H/ He] male has an even higher spontaneous liver tumor incidence than does the B6C3F1 male mouse (Graso and Hardy, 1975). This suggests the B6C3F1 mouse may have a genetic predisposition to the development of a high spontaneous liver tumor incidence. This genetic factor could be an activated cellular oncogene. 91 Table 3. Transforming genes in spontaneous hepatocellular tumors of the B5C3F1 mouse. DNA source Frequency of Transformation positive efficiency, transforming tumors foci per ug of DNA Spontaneous liver tumors 3 9 of 11 (82%) 0.005 - 0.020 (24 months) Spontaneous liver tumors b 13 of 23 (57%) 0.012 - 0.037 (24 months) a Fox and Watanabe. 1985 b Reynolds et al., 1986 92 However, when DNA isolated from the normal tissue surrounding the liver tumors was transfected into the NIH 3T3 cells no foci were detected (Fox and Watanabe, 1985). This indicates that the genetic predisposition did not involve a dominant acting oncogene. It also indicates that the oncogene detected in the spontaneous tumors was the result of a somatic event occurring within the individual liver cells. An activated H-ras oncogene was identified in the majority of the tumors that were positive in the NIH 3T3 transfection assay. The frequency of H-ras gene activation within these tumors was determined to be 66% (4 of 6) (Fox et al., 1987) and 88% (15 of 17) (Reynolds et al., 1987) in two independent studies (Table 4). This data indicates that the H-ras oncogene is frequently activated in 86C3F1 mice spontaneous liver tumors. Other dominant oncogenes apparently were also activated in the spontaneous liver tumors because not all of the NIH 3T3 cell transformants contained an activated H-ras gene. Studies by Reynolds et al. (1987) found that the raf oncogene was activated in at least one tumor and an unidentified oncogene(s) appeared to be activated in three other tumors. In addition, DNA from several spontaneous liver tumors did not cause the transformation of NIH 3T3 cells suggesting that they did not contain an activated oncogene. There are however many oncogenes which cannot be detected using the NIH 3T3 transformation system and therefore the possibility exists that one of these oncogenes may be activated in those tumors. 93 Table 4. Oncogene activation in B5C3F1 mouse liver tumors Treatment Number of Oncogene activated used to induce positive tumors tumors in NIH 3T3 assay H-ras K-ras raf unknown None a,d 23 19 0 1 3 (spontaneous) N-HO-AAF b 7 7 0 0 0 V--CARb 7 7 0 0 0 HO-DHE b 12 10 2 0 0 DMN C 14 14 0 0 0 Furfural ‘1 13 10 2 1 0 Furan d 13 9 l 0 3 a) Fox et al., 1987 b) Wiseman et al., 1986 c) Stowers et al., 1988 d) Reynolds et al., 1987 94 2.)!I'I' [ll] .1 .11.: 132E museJixemenrs To determine if and to what extent the H-ras gene is activated in chemically induced B6C3F1 mouse liver tumors, Wiseman et a1 (1986) treated 12 day old .mice with a single i.p. injection of either N—hydroxy-2- acetylaminofluorene (N -HO-AAF), vinyl carbamate (VC) or 1'-hydroxy-2',3'- dehydroestragole (HO-DHE). Multiple hepatomas developed within 7-9 months without any additional treatment. A high frequency of 3T3 cell transformation was exhibited by DNA isolated from these tumors (7 of 7 N-OH-AAF, 7of 7 VC, 11 of 11 HO-DHE) indicating the presence of an activated oncogene. The H-ras gene was identified as the transforming oncogene in the DNA from 24 of the 26 hepatomas (Table 4). Selective oligonucleotide hybridization analysis revealed that the activation of the H-ras gene in these tumors was the result of a single base pair substitution within codon 61 (Table 5). These three chemicals are potent genotoxic carcinogens whose electrophilic metabolites react with specific sites on DNA (Miller et al., 1985; Lai et al., 1985; Fennell et al., 1985). Analysis of the particular base pair substitutions occuring at codon 61 has indicated that each chemical produced a distinct pattern of H-ras mutations (Table 5). Studies in E. cali suggest that the primary DNA adduct produced by N- HO-AAF exposure is the N-(deoxyguanosin-8-yl)-aminofluorene (Lai et al., 1985). 95 Table 5. Mutation spectrum in codon 61 (CAA) of the H~ras gene in B5C3F1 mouse liver tumors. Treatment used to Number of Codon 61 mutation spectrum induce tumors tumors with codon 61 mutations AAA CGA CTA None (spontaneous) a 15 9 3 3 N-HO-AAF b 7 7 0 0 V-CAR b 7 0 1 6 HO-DHE b 10 0 5 5 DEN C 14 7 3 4 Frufural a 6 5 1 0 Furan a 5 4 1 0 a Reynolds et al., 1987 b Wiseman et al., 1986 C Stowers et al., 1988 96 This adduct should give rise to CG => A-T transversions which was the only base pair substitution observed at codon 61 in the liver tumors induced with this compound. The base substitutions detected with HO-DHE and VC were A-T => T-A transversions and AT => GC transitions at the second nucleotide position within codon 61 (Table 5). These particular nucleotide substitutions were somewhat surprising considering the types of adducts observed in other studies with these compounds. Analysis of DNA base adducts in mouse liver after a carcinogenic dose of HO-DHE revealed that greater than 90% were at the N2 position of deoxyguanosine (Fennell et al.,1985). Data for vinyl carbamate adduct formation is not available but studies with vinyl chloride, a closely related carcinogen, indicates that binding occurs primarily at the N7 position of deoxyguanosine residues (Barbin et al.,1984). Interpretation of the mutagenic specificity of these compounds is complicated by the fact that analysis is restricted to a small number of codons (12, 13, 61, 117). In addition, not all of these mutations are equally selected for since they do not have the same efficiency of inducing morphological transformation (Reynolds et al., 1987). Therefore other mutations may be present at additional sites but they are not available for analysis since they cannot be selected. Even though the specific base substitutions in the tumors were different than what would have been predicted, each of the three chemicals exhibited a unique mutational spectrum (Table 5). This suggests that these agents activated the H-ras gene as the result of a genotoxic interaction. 97 This is even more evident when these mutational spectra are compared to the mutational spectra observed in spontaneous tumors (Table 5). As can be seen in the table, the frequency distribution of codon 61 mutations of the spontaneous tumors is distinct from those of the chemically induced tumors. In a similar study, Reynolds et al. (1987) induced tumors in the B6C3F1 mouse liver by chronic exposure to either furan or furfural. As was observed by Wiseman et al. (1986) the majority of the tumors contained an activated H— ras gene (Table 4). Their results differed from that of Wiseman et al. (1986) in that not all of the activating mutations were confined to codon 61. Some of the tumors contained mutations at codons 13 and 117 (Table 6). These compounds were determined to be nonmutagenic based on the results of the Ames bacterial mutagenicity test (Reynolds et al., 1987). However, because mutations were detected at codon positions not observed in spontaneous tumors Reynolds suggests that these agents are genotoxic under in viva conditions. One of the problems with evaluating potential human risk from nongenotoxic chemicals is that they can induce animal tumors without any apparent in vitra mutagenic activity. It is important for the risk assessment process to determine with a high degree of confidence whether these agents are truly nongenotoxic in the animal. Analysis of the mutational spectrum within the H—ras gene of liver tumors induced with a particular chemical may provide a sensitive method to evaluate the presence or absence of in viva genotoxicity. 98 Table 6. Distribution of activating codon mutations within the H-ras gene from B5C3F1 mouse liver tumors. Treatment used Number of tumors Codons activated to induce analyzed tumors 61 13 117 None a ' 15 15 0 0 (spontaneous) N-HO-AAF b 7 7 0 VCAR b 7 7 0 0 HO-DHEb 10 10 0 0 DMN C_ 14 14 O O Furfural a 9 5 0 4 Furan a 9 6 2 1 3‘ Reynolds et al., 1987 b Wiseman et al., 1986 C Stowers et al., 1988 This would eliminate some of the uncertainty concerning the interpretation of these bioassays in regard to assignment of potential human risk. The basic principle underlying this type of analysis is the comparison of the mutational spectrum within the H-ras oncogene from spontaneous tumors with that of the chemically induced tumors. This comparison involves the determination in the frequency distribution of base pair substitutions within codon 61 (intracodon analysis) and also identifying which codons within the H-ras gene are activated (intercodon analysis). Statistically significant deviations, in either one of these parameters, from that observed in spontaneous tumors would be an indication of possible genotoxic activity. However, as can be seen in Table 5, diethylnitrosamine a potent genotoxic carcinogen, has a mutational spectrum in codon 61 that is very similar to that observed in spontaneous tumors. Consequently not all genotoxic agents elicit a unique mutational spectrum. If the spontaneous tumors are to be used as a benchmark, then the mutational spectrum of these tumors must be determined with a high degree of confidence. Previously only 15 spontaneous tumors have been analyzed in this manner (Reynolds et al., 1987). Analysis of additional tumors will be required to achieve the statistical confidence necessary to make the appropriate comparisons. In addition, to determine whether nongenotoxic carcinogens have cryptic mutational activity, mutational analysis of the H-ras gene in tumors induced with these types of chemicals needs to be conducted. These studies should increase the understanding of how genotoxic and 100 nongenotoxic agents activated the H—ras gene in these tumors which should aid the interpretation of bioassay data using this strain of mouse. LIST OF REFERENCES 101 References Adari, H., Lowy, D. R., Willumsen, B. M., Der, C. 1., McCormick, F., (1988), Guanosine triphosphatase activating protein (GAP) interacts with the p21 ras effector binding domain, Science, 240: 518-521. Alitalo, K., Schwab, M., Lin, C. S., Varmus, H. E., Bishop, 1. W., (1983), Homogeneously staining chromosomal regions contain amplified copies of an abundantly expressed cellular oncogene (c-myc) in malignant neuroendocrine cells from a human colon carcinoma, Proc. Natl. Acad.Sci. U.S.A., 80: 1707-1711. Andervant, H. B., (1950), Studies on the occurrence of spontaneous hepatomas in mice strains C3A and CBA,1., Natl. Cancer Inst., 11: 581-592. Angel, P., Imagawa, M., Chiu, K, Stein, B., Irnbra, R. 1., Rahmsdorf, H. 1., Ionat, C., Herrlich, P., Karin, M., (1987), Phorbol ester-inducible genes contain a common cis element recognized by a TPA-modulated trans-acting factor, Cell, 49: 729-739. Angel, P., Poting, A., Mallick, U., Rahmsdorf, H. 1., Scharpp, M. and Herrlich, P., (1986), Induction of metallothionein and other mRN A species by carcinogens on tumor promotors in primary human skin fibroblasts. Mol. Cell Biol., 6: 1760-1766. Barbicid, M., (1987), ras Genes, Annul. Rev., Biochem., 56: 779. Barbin, A., Besson, F., Perrard, M-H., Bereziat,1-C., Kaldor,1., Michel, G., Bartsch, H., (1985), Induction of specific base-pair substitutions in E. cali type A mutants by chloroethylene oxide, a carcinogenic vinyl chloride metabolite, Mutation Research, 152: 147-156. Bargmann, C. 1., Hung, M.-C., and Weinberg, R.A., (1986), Multiple independent activation of the neu oncogene by a point mutation altering the transmembrane domain of p185, Cell, 45:649-657. Becker, F. F., (1982), Morphological classification of mouse liver tumors based on biological characteristics, Cancer Research, 42: 3918-3923. Berenblum, I., (1929), Tumor formation following freezing with carbon dioxide snow, Brit. 1. Exper. Pathol., 10: 179-184. 102 Bernstein, T. Gold, L. S., Ames, B. N., Pike, M. 5., Hoel, D. G., (1985), Some tautologous aspects of the comparison of carcinogenic potency in rats and mice, Fundam. Appl. Toxicol., 5: 79-86. Bishop, 1. M., (1983), Annu. Rev. Biochem., 52: 301-354. Bishop, 1. M., The Molecular Genetics of Cancer, Science 235: 306-311. Bishop, 1.M., (1983), Cellular oncogenes and retroviruses, Annu. Rev. Biochem., 52: 301. Bister, K. and Deusberg, RH, (1982), in Advances in Viral Oncology (ed.) Klein, G. Raven Press, New York, 3-42. Bohmann, D., Bos, T. 1., Admon, A., Nishimura, T., Vogt, P. K. and T. Jan, (1987), Human proto-oncogene c-jun encodes a DNA binding protein with structural and functional properties of transcription factor AP-l, Science, 238: 1386-1392. Bonfiglio, T. A. and Stoler, M. H., (1988), Human papillomavirus and cancer of uterine cervix, Hum. Pathol. 19: 621. Bos, 1. L., (1989), ms Oncogenes in human cancer, a review, Cancer Research, 49: 4682-4689. Bos,1., (1988), The ras gene family and human carcinogenesis, Mutation Research, 195(3): 255-271. Boutwell R.K., (1964), Some biological aspects of skin carcinogenesis, Prog. Exp. Tumor Res., 4: 207. Bowen-Pope, O. F., Vogel, A., and Ross, R., (1984), Production of platelet- derived growth factor-like moleculas and reduced expression of platelet- derived growth factor receptors accompanying transformation by a wide variety of agents, Proc. Natl. Acad. Sci. U.S.A., 81: 2396-2400. Brodeur, G. M., (1986), Clinical implications of oncogene activation in human neuroblastomas, Cancer, 58: 541. Brodeur, G. M., (1986), Molecular correlates of cytogenetic abnormalities in human cancer cells: implications for oncogene activation, Prog. Hematol., 16: 229. 103 Butterworth, BE. and Slaga, T.1., (1987) Banbury Report 25: Nongenotoxic Mechanisms in Carcingenesis, Cold Spring Harbor. Campisi,1., Gray, H. E., Pardee, A. B., (1984), Cell-cycle control of c-myc but no c-r_as expression is lost following chemical transformation, Cell, 36: 241 -247. Campisi,1., Gray, H. E., Pardee, LA. B., Dean, M., Sonenshein, G. E., (1984), Cell cycle control of c-myc but not c-ras expression is lost following chemical transformation, Cell, 36: 241-247. Capon, D. 1. Chen, E. Y., Levinson, A. D., Seeburg, P. H., Goedel, D. V., (1983), Complete nucleotide sequences of T24 human bladder carcinoma oncogene and its normal homologue, Nature, 302: 33-37. Chan, C., (1984), NTP Technical Report, Cas N o. 597-25-1. Chang, E. H., Furth, M. E., Scolnick, E. M., Lowry, D. R.,(1982), Tumorigenic transformation of mammalian cells induced by a normal human gene homologous to the oncogene of Harvey murine sarcoma virus, Nature, 297: 479-483. Chang, E. H., Gonda, M. A., Ellis, R. W., Scolnich, E. M. Lowry, D. R, (1982), Human genome contains four genes homogeneous to transforming genes of Harvey and Kirsten murine sarcoma virus, Proc. Natl. Acad. Sci. U.S.A., 79: 4848-48952 Chen, Z. Q, Ulsh, L S., DuBois, G. and Shih, T., (1985), Posttranslational processing of p21 ras proteins involves palmitoylation of the c-terminal tetrapeptide containing cys 186,1. Viral, 56: 607-612. Chisari,F.V., (1989), Analysis of hepatitis B virus gene expression in transgenic mice, In Concepts in Viral Pathogenesis Vol 3, (ed.) Notkins and Oldstone, 164-169. Cichutek, K. and Duesberg, P. H., (1986), Harvey ras genes transform without codon mutations, apparently activated by truncation of a 5' exon (exon-1), Proc. Natl. Acad. Sci., 83: 2340-2344. Cifone, M. and Fidler,1., (1981), Increasing metastatic potential is associated with increasing genetic instability of clones isolated from murine neoplasma. Proc. Natl. Acad. Sci. U.S.A. 78: 6249-6252. 104 Clarke, S., Vogel, I. P., Deschenes, R. 1., Stock,1., (1988), Posttranslational modification of the H-ras oncogene protein: evidence for a third class of protein carboxyl methyltransferase, Proc. Natl. Acad. Sci. U.S.A., 85: 4643-4647. Cotran, R. 5., Kumar, V., Robbins, S. L., (1989), Neoplasia, in Pathologic Basis of Disease, W. B. Saunders Co., 239-306. Courtneidge, S. A. and Smith, A. F., (1983), Polyoma virus transforming protein associates with the product of the c-src cellular gene, Nature 303: 435- 436. Couturier, A., Bazzar, 5., and Castagna, M., (1984), Further characterization of tumor promoter mediated activation of protein kinase C, Biochem. Biophys. Res., Commun., 121: 448-455. Croce, C. M. and Nowell, P. C., (1985), Molecular basis of human B cell neoplasia. Blood 65: 1-7. Curran, T. and Franza, BR, Pas and jun: the AP-1 connection, Cell, 55: 395- 397. Curran, T., Miller, A. D., Zokas, L., Verma, I. M., (1984), Viral and cellular fos proteins: a comparative analysis, Cell, 36: 259-268. Dalla-Favera, R., Bregni, M.,. Erikson,1., Patterson, D., Gallo, R. C. and Croce, C. M., (1982), Assignment of the human c-myc oncogene to the region of chromosome 8 which is translocated in Burkett lymphoma cells, Proc. Natl. Acad. Sci. U.S.A. 79: 7824-7827. Darnell, 1., Lodish, H., and Baltimore, D., (1986), Molecular Cell Biology, W. H. Freeman and Company, N.Y., N.Y. DeCapiro, 1. A., Ludlow, 1. W., Figge,1., Shew,1-Y., Huang, C-H., Lee, W-H., Marsilio, E., Poucha, E. and Livingston, D. M., (1988), SV-40 large tumor antigen forms a specific complex with the product of the retinoblastoma susceptibility gene, Cell, 54: 275-283. Delbecco, R. and Ginsberg, H S., (1980), Virology, Harper and Row, Publisher, Philadelphia, PA. 105 Delclos, K. B., Tarpley, W., G., Miller, E. C., Miller, 1. A., (1984), 4- Aminoazobenzene and N ,N -dimethyl-4-aminoazobenzene as equipotent hepatic carcinogens in male C57BL/ 61 X C3H/He1 F1 mice and characterization of N-(deoxyguanosin-8-yl)-4-aminobenzene as the major presistent hepatic DNA-bound dye in these mice, Cancer Res. , 44: 2540-2550. DeVos, A. M., Tong, L., Milburn, M. V. Matias, P. M., 1ancarik, 1., Noguchi, S., Nishimura, S., Miura, K.,. Ohtsura, E., Kim, S-H., (1988), Three-dimensional structure of an oncogene protein: catalytic domain of human c-H-ras p21, Science, 239: 888-893. Dippold, W. G.,1ay, G., DeLev, A. B., Kharry, G., and Old, L. 1., (1981), p53 transformation-related protein detection by monoclonal antibody in mouse and human cells, Proc. Natl. Acad. Sci. U.S.A., 78: 1695-1699. Distel, R. 1., R0, H-S, Rosen, B. S., Groves, D., L., Spielgelman, B. M., (1987), Nucleoprotein complexes that regulate gene expression in adipocyte differentiation: direct participation of c-fas, Cell, 49: 35-844. Doll, R. and Hill, A. B.,(1950), Smoking and carcinoma of the lung, British Medical Journal, 2: 739-48. Doll, R., (1988), Effects of exposure to vinyl chloride, Scand. 1. Work Environ Health, 14: 61-78. Doll, R., Peto, R., (1981), The causes of cancer: quantitative estimates of avoidable risks of cancer in the United States today, 1. Natl. Cancer Inst, 66(6): 1192-1303. Downward, 1., Yarden, Y., Mayes, E., Scarce, G., Totty, N. Stockwell, P., Ulrich, A., Schlessinger,1., Waterfield, M. D., (1984), Close similarity of epidermal growth factor receptor and the v-erb B oncogene protein sequences, Nature, 307: 521-527. Drake, 1. W., Allen, E.F., Forsberg, S.A., Preparota, R.M., and Greenberg, E.O., (1969), Spontaneous mutations, Nature (London), 221: 1128-1132. Drinkwater, N. R. and Ginsler, 1. 1., (1986), Genetic control of hepatocarcinogenesis in C57BL/61 and C3H/ e1 inbred mice, Carcinogenesis, 7: 1701-17607. 106 Druckey, H., (1967), In Potential Carcinogenic Hazards from Drugs, Evaluation of Risk. Terhaut R (ed.): UTCC Monograph Series, Vol. 7. Berlin:Springer- Verley, 60-77. Duesberg, P. H. and Vogt, P. K., (1970), Differences between ribonucleic acids of transforming and nontransforming avian tumor viruses, Proc. Natl. Acad. Sci. U.S.A., 67: 1673-1680. Dutton, DR. and Bowden, G. T., (1985), Indirect induction of a clastogenic effect in epidermal cells by a tumor promoter, Carcinogenesis 6: 1279-1284. Dyson, N., Howley, P. M., Munger, K., Harlow, E., (1989), The human papilloma virus-16 E7 oncogene is able to bind to the retinoblastoma gene product, Science, 243: 934-937. Ehrlich, G. D. and Poiesy, B. 1., (1988), Clinical and molecular parameters of HTLV-l infection, Clin. Lab Med. 8: 65. Ek, B and Heldin, C-H., (1982), Characterization of a tyrosine-specific kinase activity in human fibroblast membranes stimulated by platelet-derived growth factor, 1. Biol. Chem., 257: 10486-10492. Elcomb, C. K, Rose, M. S., Pratt, I. S., (1985), Biochemical and histological changes in rat and mouse liver following the administration of tricthroethylene: possible relevance to species differences in hepatocarcinogenicity, Toxicol. and Applied Pharmacol. 79: 365-76. Environmental Health Criteria, (1975), Trichloroethylene, WHO, Geneva, 50. Erikson, 1. Nishikura, K., ar-Rushdi, A., Finan, 1., Emanuel, B., Lenoir, G., Nowell, P. and Croce, C., (1983b), Translocation of an immunoglobin kappa locus to a region 3' of an unrearranged c-myc oncogene enhances c-myc expression, Proc. Natl. Acad. Sci. USA, 80: 7581-7585. Erikson,1., Finger, 1., Sun, L., ar-Rushdi, A., Nishikura, K., Minowada, 1., Finan,1., Emanuel, B. S., Nowell, P. C., Croce, C. M., (1986), Deregulation of c- myc by translocation of the alpha-locus of the T-cell recptor in T- cell leukemias, Science, 232: 8894-886. Erikson,1., Rushdi, A., Drwinga, H. L., Nowell, P. C., Croce, C. M., (1983a), Transcriptional activation of the translocated c-myr oncogene in Burkitt's lymphoma, Proc. Natl. Acad. Sci., 80: 820-824. 107 Fasano, 1., Taparowsky, E., Fiddles, 1., Wigler, M., Goldfarb, M., (1983), Sequence and structure of the coding region of the human H-ras-l gene from T24 bladder carcinoma cells, 1. Male & Applied Gen., 2(2): 173-180. Fasano, O., Aldrich, T., Tamanoi, F., Taparowsky, E. Furth, M. and Wigler, M., (1984), Analysis of the transforming potential of the human H-ras gene by random mutagenesis, Proc. Natl.Acad. Sci., (81): 4008-4012. Fennell, T. R., Wiseman, R. W., Miller, 1. A., Miller, E., C., (1985), Major role of hepatic sulfotransferase activity in the metabolic activation, DNA adduct formation, and carcinogenicity of 1'-hydroxy-2',3'-dehydroestragole in infant male C57BL/61x C3H/ He] F1 mice, Cancer Res, 45: 5310-5320. Finlay, C., A., Hinds, P. W., Levine, A. 1., (1989), The p53 proto-oncogene can act as a suppressor of transformation, Cell, 57: 1083-1093. Food and Drug Administration, Advisory Committee on Protocols for Safety Evaluation: Panel on Carcinogenesis A Report on Cancer Testing in the Safety Evaluation of Food Additives and Pesticides, (1971), Toxicol., Appl. Pharmacol., 20: 419-438. Ford, R Maizel, A. ,(ed) (1986), Mediators in Cell Growth and Differentiation. New York: Raven Press. Fox, T. R. and Watanabe, P. G.,(1985), Detection of a cellular oncogene in spontaneous liver tumors of B5C3F1 mice, Science, 228: 596-597. Fax, T. R., Schumann, A. M., Watanabe, P., G., (1987), Activation of a cellular proto-oncogene in spontaneous liver tumor tissue of the B5C3F1 mouse, Arch. Toxicol., Suppl 10: 217-227. Franks, L. M., (1986), What is Cancer? In Cellular and Molecular Biology of Cancer Ed. L. M. Frank and N. Teich, Oxford Scientific Pub. Oxford. Frenkel, K., and Chrzan,K., (1987), DNA base modifiction by tumor promoter- activated polymorphonuclear leukoctyes, Carcinogenesis 8: 455-460. Gibbs, 1., B., Sigal, I. S., Poe, M., Scolnick, E. M., (1984), Intrinsic GTPase activity distinguishes normal and oncogenic ras p21 molecules, Proc. Natl. Acad. Sci. U.S.A., 81: 5704-5708. 108 Goldstein, B.D., Witz, G., Amoruso, M., Stone, D.S., and Troll, W., (1981), Stimulation of human polymorphonuclear leukocyte superoxide anion radical production by tumor promoters, Cancer Lett. 11: 257-262. Goustin, A. 5., Leaf, E. B., Shipley, G. D., Moses, H. L., (1986), Growth factor and cancer, Cancer Research, 46: 1015-1029. Graso, P. and Hardy, 1.,(1975), Strain differences in natural incidence and response to carcinogens, In Butler, W. H. and Newberne, P. M., (ed), Mouse Hepatic Neoplasia, Elsevier, New York, p.111-131. Grasso, P. and Hardy, 1., (1975), Strain differences in natural incidences and response to carcinogens, In Butler, W. H. and Newberne, B. M. (eds) , Mouse Hepatic Neoplasia, Elsevier, New York, pp.111-131. Groffen, 1., Stephenson, 1. R. Heisterkamp, N., deKlein, A., Bartram, C. and Grosveld, G., (1984), Philadelphia chromosome breakpoints are clustered within a limited region, bcr, on chromosome 22, Cell 36: 93-99. Gross, L., (1983), Oncogenic viruses, 3rd ed., Pergamon Press, NY., NY. Gutierres, L. Magee, A. 1., Marshall, C. 1., and Hancock, 1. F., (1989), Post- translational processing of p21 ras is a two-step and involves carboxyl- methylation and carboxy-terminal proteolysis, EMBO], 8: 1093-1088. Hamm, T. E., (1984), Occurrence of Mouse Liver Neoplasma in the National Cancer Institute Bioassays ed. Popps 1. A. , Hemisphere Publishing, Mouse Liver Neoplasma, Current Perspective, p.27-38. Hancock, 1. F., Magee, A. 1., Childs, 1. E. and Marshall, C. 1., (1989), All ras proteins are polyisoprenylated but only some are palmitolyated, Cell, 57: 1167- 1172. Hanigan, M., H., Kemp, C. 1., Ginsler,1., 1. and Drinkwater, N. R., (1988), Rapid growth of preneoplastic lesions in hepatocarcinogen-sensitive C3H/He1 male mice relative to C57BL/61 male mice, Carcinogenesis, 9(6): 885-890. Hansen, M., and Cavenee, W., (1988), Tumor suppressors: Recessive mutations that lead to cancer, Cell, 53: 172-173. 109 Harris, 1. F., Chambers, A. F., Hill, R. P. and Ling, V., (1981), Metastatic varients are generated spontaneously at a high rate in mouse KI-IT tumor, Proc. Natl. Acad. Sci. U.S.A. 79: 5547-5551. Hartman, P. E., (1980), Bacterial mutagenesis - review of new insights, Environ. Mut., 2: 3-16. Heisterkamp, N., Stephenson, 1. R., Goffen, 1., Hansen, P. F., deKlein, A., Bartram, C. R., Grosveld, G., ( ), Localization of the c-abl oncogene adjacent to a translocation breakpoint in chronic myelocytic leukemia, Nature, 306: 239- 242. Heppner, G. H., (1984), Tumor heterogeneity, Cancer Research, 44: 2259-2265. Hill, D. E., Hope, I.A., Macke, 1. P., Struhl, K., (1986), Saturation mutagenesis of the yeast his3 regulatory site: requirements for transcriptional induction and for binding by GCN 4 activator protein, Science, 234: 451-457. Hollis, G. F., Mitlchell, K. F., Battey,1., Potter, H., Taub, R. A., Lenoir, G. Ledey, P., (1984), A varient translocation places the lambda immunoglobin genes 3' to the c-myg oncogene in Burkett lymphoma, Nature, 307:752-755. Hope, I. A., Struhl, F.,(1985), GCN4 protein, synthesized in vitro, binds HIS3 regulatory sequences: implications for general control of amino acid biosynthetic gene in yeast, Cell, 43: 177-188. Houwelling, A., Van der Elsen, P. 1. and Van der Eb, A. 1., (1980), Partial transformation of primary rat cells by the leftmost 4.5% fragment of adenovirus 5 DNA, Virology, 105: 537-550. Huang, 1. 5., Huang, 5. S. and Deuel, T. F., (1984), Transforming protein of simian sarcoma virus stimulates autocrine grthh of SSV - transformed cells through PDGF cell - surface receptors, Cell, 39: 79-87. Huebner, RJ. and Todaro, G.1., (1969), Oncogenes of RNA tumor viruses as determinants of cancer, Proc. Natl. Acad. Sci. U.S.A., 64: 1087. Hunter, T., (1984), Oncogenes and proto-oncogenes: How do they differ, 1NCI, 73(4): 773-786. 110 Hurley, 1. B., Simon, M. I., Teplow, D. B. Robishaw, J. D., Gilman, A. G., (1984). Homologies between signal transducing G proteins and ras gene products, Science, 226: 860-862. Hurlin, P. 1., Fry, D. G., Maher, V. M., McCormick, J. J., (1987), Morphological transformation, focus formation and anchorage independence induced in diploid human fibroblasts by expression of a transfected H-ras, Cancer Res., 47: 5752-5757. Hurlin, P. 1., Maher, V. M., McCormick, J. 1., (1989), Malignant transformation of human fibroblasts cause by expression of a transfected T24 H-ras oncogene, P.N.A.S., 86: 187-191. IARC Monograph, (1980), Long-term and short-term assays for screening carcinogens: a critical appraisal, Supplement #2, Interdisciplinary Panel on Carcinogenicity, (1984), Criteria for evidence of chemical carcinogenicity, Science, 225: 682. Inversen, U. M. and Inversen, O. H., (1979), The carcinogenic effect of TPA when applied to the skin of hairless mice, Virchows Arch. 30:33-42. Jensen, R. K., Sleight, S. D., Goodman, J. I., Aust, S. D., and Trosko, J. E. (1982), Polybrominated biphenyls as promoters in experimental hepatocarcinogensis in rats, Carcinogenesis, 3: 1183-1186. Jochemsen, H., Daniels, G.S.G., Hertoghs, P.1., Schrier, van den Elsen, P.1., and van der Eb, A.1., 1982, Identification of adenovirus type 12 gene products involved in transformation and oncogenesis, Virology, 122: 15 Johnson, A., Heldin, C-H, Westermark, G., Wasteson, A., (1982), Platelet- derived growth factor: identification of constituent polypeptide chains, Biochem. Biophys. Res. Comm., 104: 66-74. Kakunaga, T. and Crow, 1. D., (1980), Cell varients showing differential susceptibility to ultraviolet light-induced transformation, Science 209: 505-507. Kaminski, M. and Auerbach, R., (1988), Tumor cells are protected from NK cell-medicated lysis by adhesion to endothelial cells, Int. J. Cancer 41: 847. Klein, G.‘, (1983), Specific chromosomal translocation and the genesis of B-cell- derived tumors in mice and men, Cell, 32: 311-315. 111 Klein, G., (1987), The approaching era of the tumor suppressor genes, Nature, 238: 1539-1545. Klinger, H. P., (1982), Suppression of tumorigenicity, Cytogenet. and Cell. Genet. , 32: 68-84.‘ Kluwe, W. M., Huff, J. E., Matthews, H. B., Irwin, R., Haseman, J. K., (1985), Comparative chronic toxicities and carcinogenic potentials of 2-ethylhexyl- containing compounds in rats and mice, Carcinogenesis 6: 1577-83. Knudson, A. G., (1971), Mutation and cancer: statistical study of retinoblastoma, Proc. Natl. Acad. Sci. U.S.A, 68: 820-823. Krieg, P., Finch, J., Fustenberger, G., Melber, K., Matrisian, L. M., Bowden, G. T., (1988), Tumor promoters induce a transient expression of tumor- associated genes in both basel and differentiated cells of the mouse epidermis, Carcinogenesis, 9 (1): 95-100. Lacal,1. C., Srivastava, S. K., Anderson, P. 5., Aaronson, S. A., (1986), Ras p21 proteins with high or low GTPase activity can efficiently transform NIH/3T3 cells, Cell, 44: 609-617. Lai, C-C, Miller, J. A., Miller, E. C., Liem, A., (1985), N sulfoxy-2- animofluorene is the major ultimate electrophilic and carcinogenic metabolite of N hydroxy-2—acetylaminofluorene in the livers of infant male C57Bl / 61 x C3H/ HeJ F1 B6C3F1 mice, Carcinogenesis 6 (7): 1037-1046. Land, H., Parada, L. F. and Weinberg, R. A., (1983), Tumorigenic conversion of primary embryo fibroblasts requires at least two cooperating oncogenes, Nature, 304: 596-602. Land, H., Parada, L., and Weinberg, R. A., (1983), Cellular oncogenes and multistep carcinogenesis, Science, 222: 771 -778. Lane, D. P. and Crawford, L. V., (1979), T Antigen is bound to a host protein in SV40 transformed cells, Nature, 278: 261-263. Lee, W. Mitchell, D., Tjian, R., (1987), Purified transcription factor AP—l interacts with TPA-inducible enhancer elements, Cell, 49: 741-752. 112 Ling, V., Chambers, A. F., Harris, J. F. and Hill, R. P., (1985.), Quantitative genetic analysis of tumor progression, Cancer Metastasis Rev., 4: 173-194. Liotta, L. A. et al., (1988), Biochemical mechanisms of tumor invasion and metastases, Prog. Clin. Biol. Res. 256: 3. Lippman, M. F., dickson,R.B., Gelmann, E.P., Rosen, N., Kaufman, D., Knabbe, C., Bates, S., Kasid, A., Salomon, D., (1987), Growth regulation of human breast carcinoma occurs through regulated growth factor secretion, 1. Cell Biochem. 35 (1): 1-16. Loeb, L. A., Weymouth, L. A., Kunkel, T. A., Gopinathan, K. P., Beckman, R. A., Dube, D. F., (1978), Cold Spring Harbor Symposium Quant., Biol. 43: 921- 927. Loury, DJ. Goldsworthy, T.L., and Butterworth, BE, (1987), The role of measuring cell replication as a predictive index of tissue-specific tumorigenic potential, Banbury Report 25: Nongenotoxic Mechanisms in Carcingenesis, Cold Spring Harbor. MacMahon, B., (1979), Epidemiology's strengths, weaknesses in environmental health issues, The National Research Council in 1979: Current Issues and Studies, p340. Maher,1. M.,, Dorney, D. 1., Mendrala, A. L, Konge-Thomas, B., McCormick, J. 1., (1979), DNA excission-repair processes in human cells can eliminate the cytotoxic and mutagenic consequences of ultraviolet irradation, Mutation Research, 62: 311-323. Maki, Y., Bos, T. 1., Davis, C., Starbuck, M. land Vogt, P. K., (1987), Avain sarcoma virus 17 carries the jun oncogene, Proc. Natl. Acad. Sci., 84: 284-2852. Manolova, G. and Manolova, Y., (1972), Marker band in one choromosome 14 from Burkitt lymphoma, Nature 237: 33-34. Manne, V., Bekesi, E. and Kung, H., (1985), H-ras proteins exhibit GTPase activity: point mutations that activate Ha-ras gene products result in decreased GTPase activity, Proc. Natl. Acad.Sci U.S.A., 82: 376-380. Marx, J. L. , How DNA viruses may cause cancer. (1989), Science, 243: 1012- 1013. 113 McCann, H. 1.., Choi, E. Yamasaki, E., Ames, B. N., (1975), Detection of carcinogens as mutagens in the Salmonella / microsomal test: assay of 300 chemicals, Proc. Natl. Acad. Sci. U.S.A., 72(12): 5135-5139. McCann, J. Choi, E., Yamasaki, Ames B., (1975), Detection of carcinogens as mutagens in the Salmonella microsomal test: Assay of 300 chemicals, Proc. Natl. Acad. Sci. U.S.A., Vol. 72: 135-5139. McCormick, F.,(1989), ras GTPase activating protein: Signal transmitter and signal terminator, Cell, 56: 5-8. McGrath ,1. P., Capon, D. J., Goedel, D. V., Levinson, A. D.,(1984), Comparative biochemical properties of normal and activated human ras p21 protein, Nature, 310: 644-649. McGrath, J. F., Capon, D. J., Smith, D. H., Chen, E. Y., Seeburg, P. H., Goedel, D. V. and Levinson, A. D., (1983), Structure and organization of the human K- ras proto-oncogene and a related processed pseudogene, Nature, 304: 501-506. Miller, E. C. and Miller, J. A., (1981), Mechanisms of chemical carcinogenesis, Cancer 47:1055—1064. Miller, E. C., Miller, J. A., Boberg, E. W., Delclos, K. B., Lai, C-C, Fennell, T. R., Wiseman, R. W., Liem, A., (1985), Sulfuric acid esters as ultimate electrophilic and carcinogenic metabolites of some alkenylbenzenes and aromatic amines in mouse liver, Carcinogenesis, 10: 93-107. Mirsalis,1. C., Tyson, C. K., Loh, E. N., Steinmetz, K. L., Bakke, J. P., Hamilton, C. M., Spak, D. K. and Spalding , J. W., (1985), Induction of hepatic cell proliferation and unscheduled DNA synthesis in mouse hepatocytes folloWingin viva treatment, Carcinogenesis, 6: 1521-1524. Mougneau, E., Lemieux, L., Russoulzadegan, M., Cuzin, F., (1984), Biological activities of v-myc and rearranged c-myc oncogenes in rat fibroblast cells in culture, Proc. Natl. Acad. Sci., 81: 5758-5762. Muir, C. S.,(1979), Limitations and advantages of epidemiological investigations in environmental carcinogenesis (1979), Annals New York Academy of Sciences, 153-165. 114 Nerenberg, M., Hinrechs, 5. H., Reynolds, R.K., Khoury, G., Jay, G., (1987), The fat gene of human T-lymphototropic virus type 1 induces mesenchymal tumors in transgenic mice, Science 237: 1324-1329. Nicolson, G. and Milos, L., (ed). (1984), Cancer Invasion and Metastasis: Biologic and Therapeutic Aspects, New York, Raven Press. Nishimura, 5., and Sekiya, T., (1987), Human cancer and cellular oncogenes, J. of Biochem., 243: 313-327. Nowell, P. C., (1986), Mechanism of Tumor Progression, Cancer Research 46: 2203-2207. Nutrition Foundation, (1983), The relevance of mouse liver hepatomas to humand carcinogenic risk, Report to the international expert advisory committee to the Nutrition Foundation, The Nutrition Foundation, Inc, Washinton, D.C. O'Connell, J. F., Rotstein, J.B., Slaga, T.1., (1987), Tumor promotion and progression in the mouse skin model, In Banbury Report 25: Nongenotoxic Mechanisms in Carcingenesis, Cold Spring Harbor, 11-24. Oksala, T. and Therman, E., (1979), Mitotic abnormalities in cancer. Chromosomes and Cancer. 1. German (ed) pp. 239-263, New York. John Wiley 8: Sons . Omar, R. A. and Lanks, K. W., (1983), Origin of phenotypic variation in clones of smian virus 40-transformed mouse embryo cells, Cancer Res., 43: 1835-1841. Oren, M., Maltgman, W., and Levine, A. 1., (1981), Post-translational regulation of 54 K cellular tumor antigen in normal and transformed cells, Mol. Cell. Biol., 1: 101-110. 'Pacifici, G. M., Boobis, A. R., Brodie, M. 1., McManus, M. E., Davis, D. 5., (1981), Tissue and species differences in enzymes of epoxide metabolism, Xenobiotica, 11: 73-79. Persson, H. and Iede, D., (1984), Nuclear localization and DNA binding properties expressed by human c-myc oncogene, Science, 225: 718-721. 115 Peterson, 1. A., Ceriani, R. L., Blank, E. W. and Osvaldo, L., (1983), Comparison of rates of phenotypic variability in surface antigen expression in normal and cancerous human breast epithelial cells, Cancer Research 43: 4291-4296. Pitot, H. C. Goldsworthy, T., Cambell, H. A., Poland, A., ( 1980 b), Quantitative evaluation of the promotion by 2, 3, 7 8-tetrachlorodibenzo-p-dioxin of hpatocarcinogenesis for diethylnitrosamine, Cancer Research 40 (10): 3616. Pitot, H. C., and Sirica, A.E., (1980 a), The stages of initiation and promotion in hepatocarogenesis, Biochem. Biophys. Acta, 605: 191-215. . Pitot, H. C., Goldsworthy, T., Campbell, H. A. and Poland, A., (1980), Quantitative evaluation of the promotion by 2,3,7,8-tetrachlorodibenzo-p- dioxin of hepatocarcinogenesis from diethylnitrosamine, Cancer Res, 40: 3616-3620. Pitot, H. C., Goldworthy, T. and Moran, S., (1982), The natural history of carcinogenesis: Implications of experimental carcinogenesis in the genesis of human cancer, Mechanism of Chemical Carcinogenesis. Ed. Harris, C. C., Cerutti, P. A. Pub. Alan R Liss Inc., NY, NY. Pitot, H.C., (1986) Fundamentals of Oncology, 3rd edition, Marcel Decker, New York. ' Piwnica-Worms, H., Saunders, K. B., Roberts, T. M., Smith, A. E., Cheng, S. H., (1987), Tyrosine phosphorylation regulates the biochemical and biological properties of pp60 c-src, Cell, 49: 75-82. Poland, A. and Knutson, J. C., (1982), 2,3,7,8-Tetrachlorodibenzo-p-dioxin and related aromatic hydrocarbons: examination of the mechanism of toxicity, Ann. Rev. Pharmacol. 22: 217-54. Poland, A., Glover, E., (1975), Genetic expression of aryl hydrocarbon hydroxylase by 2, 3, 7, 8-tetrachlorodibenzo-p-dioxin evidence for a receptor mutation in genetically nonresponsive mice, Mol. Pharmacol. 11(4): 389-398. Ponticelli, A. 5., Whitlock, C. A., Rosenberg, N. and Witte, ON, (1982), In viva tyrosine phosphorylation of the Abelson virus transforming protein are absent in its normal cell homolog, Cell, 29: 953-960. 116 Powers, 5. Katovka, T., Fasano, O., Goldfarb, M., Stratherm,1., Broach, 1. and Wigler, M., (1984), Genes in S. cerevisiae encoding proteins with domains homologous to the mammalian ras protein, Cell, 36: 607-612. Pozzati, R, Muschel, R, Williams, 1., Padmanabhan, R., Howard, B., Liotta, L., Khoury, G., (1986), Primary rat embryo transformed by one or two oncogenes show different, metastatic potentials, Science, 232: 223-227. Pulciani, 5., Santos, E., Lauver, A. V., Long, L. K., Aaronson, S. A., Barbacid, M.,(1982), Oncogenes in solid human tumors, Nature, 300: 539-542. Rassoulzadegan, M., Cowie, A., Carr, A., Glaichenhous, N., Kamen, R., Cuzin, F., (1982),The roles of individual polyoma virus early proteins in oncogenic transformation, Nature, 300: 713-718. Reddy, E., P., Lipman, D., Andersen, P. R, Tronick, 5. R., Aaronson, S. A., (1985), Nucleotide sequence analysis of the balb/ c murine sarcoma virus transforming gene, J. Virol. 53: 984-987. Reddy, 1. K. and Lalwani, N. D., (1983), Carcinogenesis by hepatic peroxisomal proliferations: evaluation of the risk of hypolipidemic drugs and industrial plasticizers to humans, CRC Crit. Rev. Toxicol. 12: 1-58. Reddy, J.K., and Lalwani, ND, (1983), Carcinogenesis by hepatic peroxisome porliferatiors: Evaluation of the fisk of hypolipidemic drugs and industrial plasticizers to humans, CRC Crit. Rev. Toxicol. 12: 1-58. Reichert, D., (1983), Biolgical actions and interactions of tetrachloroethylene, Mutation Res, 123: 411-429. Reynolds, 5. H., Stowers, S. 1., Maronpot, R. R., Anderson, I. M.. W., Aaronson, S. A.,(1986), Detection and identification of activated oncogenes in spontaneously occurring benign and malignant hepatocellular tumors of the B6C3F1 mouse, Proc. Natl. Acad. Sci. U.S.A., 83: 33-37. Reynolds, 5. H., Stowers, 5. 1., Patterson, R M., Maronpot, R R, Aaronson, S. A. and Anderson, M. W., (1987) Activated oncogenes in B5C3F1 mouse liver tumors: Implications for risk assessment. Science, 237: 1309-1316 Reynolds, 5. H., Stowers, S. 1., Patterson, R. M., Maronpot, R. R, Aaronson, S. A., Anderson, M. W.,(1987), Activated oncogenes in B6C3F1 mouse liver tumors; Implications for risk assessment, Science 237: 1309-1316. 117 Ruby, H. E., (1983), Adenovirus early region 1A enables viral and cellular transforming genes to transform primary cells in culture, Nature, 304: 602- 606. Ruley, H. E.,(1983), Adenovirus early region 1A enables viral and cellular transforming genes to transform primary cells in culture, Nature, 304: 602- 606. - Sainsbury, J.RC., (1988), Epidermal growth factor receptor status as a predictor of early recurrence of and death from breast cancer, Lancet, 1: 1398. Santos, E. and Nebreda, A. R, (1989), Structural and functional properties of ras proteins, The FASEB Journal, 3(10): 2151-2163. Sarnow, P., Ho, Y. S. , William, J., and Levine, A. J., (1982), Adenovirus E1b-58 kd tumor antigen and SV-40 large tumor antigen are physically associated with the same 54 kd cellular protein in transformed cells, Cell, 28: 387-394. Sassone-Corsi, P., Lamph, W. W., Kamps, M., Verma, I. M. (1988), fas- Associated cellular p39 is related to nuclear transcription factor AP-l, Cell, 54: 553-560. Schecter, A. L., Hung, M. C., Vardyanathan, L., Decker, S. 1., Drebin, J. A., Greene, M. I. and Weinberg, R. A., (1984), The neu oncogene: an erb b-related gene encoding a 185,000 Mr tumor antigen, Nature, 312: 513-516. Schulte-Hermann, R, Schuppler, 1., Ohde, G., and Timmermann-Trosiemer, I., (1982), Effect of tumor promoters on proliferation of putative preneoplastic cells in rat liver, E. Hecker, N. Fusenig, W. K. Lung, F. Marks and H. Thielman (eds),‘In Carcinogenesis, Vol. 7, Raven Press, New York, pp99-104 (1982). Schulte-Hermann, R., Timmermann-Trosiener, I. and Schuppler, J., (1983), Promotion of spontaneous preneoplastic cells in rat liver as a possible explaination of tumor production by nonmutagenic compounds, Cancer Res, 43:839-844. Schwab, M., Alitalo, K., Varmus, H. E., Bishop, 1. M., (1983), A cellular oncogene (c-Ki-ras) is ampliphied, overexpressed, and located within karyotypic abnormalities in mouse adrenocortical tumour cells, Nature, 303: 497-501. 118 Scolnick, F. M., Papogeorge, A. G., Shih, T. Y., (1979), Guanine nucleotide- binding activity as an assay for src protein of rat-derived murine sarcoma virus, Proc. Natl. Acad. Sci., 76: 5355-5359. Seeburg, P. H. Colby, W. W., Capon, D. 1., Goedel, D. V. and Levinson, A. D., (1984), Biological properties of human c-Ha-ras 1 genes mutated at codon 12, Nature, 312: 71-75. Sher, C. 1., Rettenmier, C. W., Socca, R, Roussel, M. F., Look, A. T. Stanley, E., R, (1985), The c-fms proto—oncogene product is related to the receptor for the mononuclear phagocyte growth factor CSF-l, Cell, 41: 665-676. Shilo, B-Z., Weinberg, R A., (1981), DNA sequences homologous to vertebrate oncogenes are conserved in Qrgsophila melanogasler, Proc. Natl. Acad. Sci. U.S.A., 78: 6789-6792. Shimizu, K., Goldfarb, M., Suard, Y., Perucha, M., Yen, L., Kamota, T., Feramisco,1., Stavnezer, E., Fogh, 1., Wigler, M. H., (1983), Three human transforming genes are related to the viral ras oncogenes, Proc. Natl .Acad. Sci., 80: 2112-2116. Slamon, D. 1., (1987), Human breast cancer: Correlation of relapse and survival with amplification of the HER-Z/neu oncogene, Science, 235: 177. Smets, L. A., (1980), Cell transformation as a model for tumor induction and neoplastic growth, Biochem. Biophys. Acta, 605: 93-111. Srnit, V. T., Boot, A. J. M, Smits, A. M.M, Fleuren, G. 1., Comelisse, C. 1., and Bos, J. L., (1988), K-ras codon 12 mutations occur very frequently in pancreatic adenocarcinoma, Nuclei Acid, Res, 16: 7773-7782. Solt, D., Farber, E., (1975), New principle for the analysis of chemical carcinogenesis, Nature, 263(5579): 701-703. Spandidos, D. A. and Wilkie, N. M., (1984), Malignant transformation of early passage rodent cells by a single mutated human oncogene, Nature, 310: 469- 475. SpandidOs, D. A., and Kerr, 1. B., (1984), Elevated expression of the human ras oncogene family in premalignant and malignant tumour of the colorectum, Br. J. Cancer, 49: 681-688. 119 Sporn, M. and Todaro, G., (1980), Autocrine secretion and malignant transformation of cells, N. Engl. J. Med., 303: 878-878. Srinivasan, A., Peden, K. W. C. and Pipas, 1. M., (1989), The large tumor antigen of simian virus 40 encodes at least two distinct transforming functions, 1. of Virology, 63 (12): 5459-5463. Stewart, T. A., Pattengale, P. K, Ieder, P., (1984), Spontaneous mammary adenocarcinoma in transgenic mice that carry and express MTV / myc fusion genes, Cell, 38: 627-637. Stowers, S. 1., Wiseman, R. W., Ward, J. M., Miller, E. C. Miller, J. A., Anderson M. W., Eva, A., (1988), Detection of activated proto-oncogenes in N- nitrosodiethylamine induced liver tumors: a comparison between B6C3F1 mice and Fischer 344 rats, Carcinogenesis, 9 (2): 271-276. Takeya, T., Hanafusa, H., (1983), Structure and sequence of the cellular gene homologous to the RSV src gene and the mechanism for generating the transforming virus, Cell, 32: 881-890. Task Force of Past Presidents - Society of Toxicology, (1982), Fundamental and App] Tox., 2: 101. ' Taub, R, Kirsch, I., Morton, C., Lenoir, G., Swan, D., Tronick, 5., Aaronson, 5., and Leder, P., (1982), Translocation of the c-myc gene into the immunoglobin heavy chain locus in human Burkett's lymphoma and murine plasmacytoma cells, Proc. Natl. Acad. Sci. U.S.A., 79: 7837-784. Teich, M. N ., (1986) In Cellular and Molecular Biology of Cancer, Ed. Frank, L.M. and Teich, N., Oxford Science Pub. Trahey, MN. and McCormick, F., (1987), A cytoplasmic protein stimulates normal N-ras p21 GTPase, but does not affect oncogenic mutants, Science, 238: 542-545. Trahey, MN ., Milley, R. 1., Cole, G. E., Innis, M., Paterson, H., Marshall, C. 1., Hall, McCormick, F., (1987a), Biochemical and biological properties of the human N-ras p21 protein, Mol. Cell Biol., 7(1): 541-544. 120 United States Public Health Service, (1989), The health consequences of smoking - Cancer: A report of the surgeon general, Rockville, MD., United States Dept. of Health and Human Services, Office on Smoking and Health. Van Beveren, C., van Stroaten, F., Curran, T., Muller, R, Verma, I. M., (1983), Analysis of FBT-MuSV provirus and c- fos (mouse) gene reveals that viral and cellular fos gene products have different carboxy termini, Cell, 32: 1241- 1255. van dem Berhe, H., Parloir, C. Gosseyl, 5., Eglebienne,1., Comu, G. and Sokal, G., (1976), Varient translocations in Burkett lymphoma, Cancer Genet. Cytogenet., 1: 9-14. Verma, A. K.,. Boutwell, R K., (1980), Effects of dose and duration of treatment with the tumor promoting agent 12-O- tetra decanoyl phorbol 13 acetate on mouse skin carcinogenesis, Carcinogenesis, 1: 271-276. Vogt, P. F., Bos, T. J. and Doolittle, R. F., (1987), Homology between the DNA binding domain and the GCN 4 regulatory protein of yeast and the carboxyl- terrninal region of a protein coded for by the oncogene jun, Proc. Natl. Acad. Sci. U.S.A, 84: 3316-3319. Wanghorne, C., Thomas, M., Lagarde, A., Kerbel, R. 5., Breitrnan, M. L., (1988), Genetic evidence for progressive selection and overgrowth of primary tumors by metatstatic cell subpopulations, Cancer Res, 48 (21): 6109-6114. Waterfield, M. D., (1986), in Cellular Molecular Biology of Cancer, Ed. Frank L. M. and Teich, N., Oxford Science Pub. ‘ Waterfield, M. D., Scarce, G. T., Whittle, N. Stroobant, P., Johnson, A., Wasteson, A. Westermark, B., Heldin, C. H., Huang, 1. 5., Deuel, T. F., (1983), Platelet-derived growth factor is structurally related to the potential transforming protein p28 sis of simian sarcoma virus, Nature, 304: 35-39. WHO (192), IARC, (1982), Monograph on the evaluation of the carcinogenic risk of chemicals to humans, Supplement 4. WHO, (1969), Principles for the Testing and Evaluation of Drugs for Carcinogenicity , WHO Tech. Report N o. 426, Geneva, World Health Organization 121 WHO, (1978), Carcinogenicity and mutagenicity, In Principles and Methods for Evaluating the Toxicity of Chemicals Part 1 Geneva, World Health Organization p. 236-272. Whyte, P., Buchkovich, K. 1., Horowitz, J. M., Friend, S.H., Raybuck, M., Weinberg, R A., Harlow, E., (1988), Association between an oncogene and an anti-oncogene: the adenovirus E1A proteins bind to the retinoblastoma gene product, Nature, 334: 124-129. Wigley, C., (1986), Chemical Carcinogenesis, In Cellular and Molecular Biology of Cancer. ed. L. M. Franks and N. Teich, Oxford Scientific Pub., Oxford ' Williams G. and 'Weisburger, (1986), Chemical Carcinogens in Toxicology, The Basic Science of Poisons, ed. Casarett and Doull, Macmillan Publishing Co., NY. Williams, G. M., Kroes, R, Waayers, H. W., van de Poll, K. W., (1980), The predictive 'value of short-term screening tests in carcinogenicity evaluation, Applied Methods in Oncology, Vol. 3. Willumsen, B. M., Norris, K., Papageorge, A. G., Hubbert, N. L. and Lowry, D., R, (1984), Harvey murine sarcoma virus p21 ras protein: biological and biochemical significance of the cysteine nearest the carboxy terminus, J. EMBO, 3: 2581-2585. Wilson, D. M., Fry, D.G., Maher, V. M., McCormick, J. 1., (1989), Transformation of diploid human fibroblasts by transfection of N-ras oncogenes, Carcinogenesis, 10(4): 635—640. Wiseman, R. W., Stowers, S. 1., Miller, E. C., Anderson, M. W. and Miller,J.A., Activating mutations of the C-Ha-ras proto-oncogene in chemically induced hepatomas of the male B6C3F1 mouse. Eggs. NsLl. Acad. &- USA, 83: 5825- 5829, 1986. Yamagiwa, K. and Ichikawa, K., (1918), Experimental study of the pathogenesis of carcinoma, 1. Cancer Res. 3: 1-29. Yunes, J. J. and Ramsey, N., (1979), Retinoblastoma and subband deletion of chromosome 13, Am. J. Dis. Child, 132: 161-163. 122 CHAPTER II. MUTATIONAL ANALYSIS or THE H-ras ONCOCENE lN ' SPONTANEOUS 35cm MOUSE LIVER TUMORS AND TUMORS INDUCED WITH A GENOTOXIC AND NONGENOTOXIC HEPATOCARCINOGENS Tony R Foxarb, Alan M. Schumanna, Philip G. Watanabea, Barry L. Yanoa, Veronica M. Maherb, 1. Justin McCormickb aThe Toxicology Research Laboratory, The Dow Chemical Company, Midland, Michigan 48674. bCarcinogenesis Laboratory, Fee Hall, Michigan State University, East Lansing, NH 48824-1316. 123 MACE The frequency and mutational profile of. H-ras gene activation was determined in spontaneous liver tumors of male B6C3F1 mice and in tumors induced with the genotoxic hepatocarcinogen benzidine-2H0 or the nongenotoxic hepatocarcinogens phenobarbital, chloroform or ciprofibrate. DNA sequence analysis of the H-ras gene from representative tumors revealed that 32 of 50 (64%) spontaneous tumors and 13 of 22 (59%) benzidine-ZHCl induced tumors contained a point mutation in codon 61. Tumors induced with the nongenotoxic agents had a much lower frequency of codon 61 mutations, i.e., phenobarbital 1 of 15 (7%), chloroform 5 of 24 (21%) and ciprofibrate 8 of 39 (21 %). No mutations were observed at codons 12, 13 or 117 in tumors from any of the groups. Only three base pair substitutions within codon 61 were found. The one most frequently detected in all of the groups was a CC to AT transversion at the lst nucleotide position, occurring at a 59%, 85%, 100%, 80%, and 88%, frequency in the spontaneous tumors and in the tumors induced with benzidine 2-HC1, phenobarbital, chloroform and ciprofibrate respectively. In these same groups a ArT to CC transition or a ArT to T-A transversion at the second nucleotide position occurred at a frequency of 34%, 8%, 0%, 0%, 12% and 6%, 8%, 0%, 20%, 0% respectively. The number of tumors carrying an activated H-ras gene in the nongenotoxic treatment groups is within the range that would be expected if those animals had not received any treatment. This indicates that the activation of the H-ras gene in those tumors is probably the result of a 124 spontaneous event. The data suggest that these toxicologically and pharmacologically diverse nongenotoxic hepatocarcinogens increase the frequency of liver tumors but do not induce mutations in the H-ras gene. Instead these agents appear to interact with a population of cells that do not contain an activated H-ras gene. This suggests that the mechanisms of tumor development by these nongenotoxic carcinogens differs at least partially from the mechanisms responsible for the development of spontaneous tumors or those induced by a typical genotoxic agent. W The results of the rodent bioassay play an important role in the evaluation of whether a chemical is considered potentially carcinogenic for humans. However, the use of such test results to predict human risk is often complicated by biological factors peculiar to the strain of animal used. For example, the B6C3F1 mouse is frequently used to assay various potential carcinogens. Male B5C3F1 mice have an average spontaneous liver tumor incidence of 20-30% (1). These mice are apparently hypersensitive to the development of such tumors which calls into question the relevance of a chemically induced increase in tumor frequency as a predictor of human cancer risk. This concern is underscored by the fact that many genotoxic and nongenotoxic agents significantly increase the frequency of liver tumors in this animal but may not be tumorigenic at other sites or in other species (2). 125 For example, Ashby and Tennant surveyed 222 carcinogen bioassays conducted by the NCI/NT'P (3). Of the 115 chemicals found to be tumorigenic, nearly 40% were hepatocarcinogenic in the B6C 3F 1 mouse . For 20% of the 1 15 carcinogens the only tumorigenic response was an increase in mouse liver tumors. Based on in m mutagenicity tests many of these tumorigenic chemicals have been shown to have genotoxic activity (ie. damage DNA). The mutagenic activity of these agents can be initiated from the parent molecule or from electrophilic metabolic intermediates. It is believed that damage to DNA resulting in somatic mutations plays an important role in the tumorigenicity of this class of chemical carcinogens (4). Not all carcinogenic chemicals have mutagenic activity as determined by in vitra mutagenicity assays. An important class of chemical carcinogens, generally referred to as nongenotoxic, do not cause mutations but still can cause an increase in animal tumors. The mechanism by which these agents increase mouse liver tumors is presently unclear. They may exert their carcinogenic activity by promoting previously inititated cells (5). Alternatively, these agents may induce somatic mutations by indirectly damaging DNA as a consequence to some abnormally induced physiologic response. For example, this secondary damage to DNA may occur as a result of the interaction of reactive oxygen radicals with DNA generated by the infiltration of polymorphonuclear leukocytes as a consequence to chemically induced cellular toxicity (6). Much of the controversy surrounding the interpretation of bioassay results concerns the relevance to human risk from exposure to nongenotoxic 126 agents that cause an increase in mouse liver tumors. In an attempt to gain a better understanding of how liver tumors arise in the B6C3F1 mouse and how this tumor incidence is increased by chemical exposure, several groups have examined the role of cellular oncogenes in the hepatocarcinogenic process of the B6C3F1 mouse. It was demonstrated. that the H-ras oncogene is activated in a high percentage of both spontaneous tumors and tumors induced with several genotoxic hepatocarcinogens (7-10). In addition Wiseman et al (9) and Reynolds et al (10) showed that certain genotoxic carcinogens cause chemical specific mutational patterns within the H-ras gene. To investigate the mechanisms by which nongenotoxic hepatocarcinogens enhance liver tumor development in the B6C3F1 mouse, we determined the frequency and mutational profile of H-ras gene activation in tumors that arose spontaneously and in tumors induced by exposure to a genotoxic or several nongenotoxic carcinogens. This was carried out by amplifying regions of exons 1, 2, and 3 of the H-ras gene which include the activating codons 12, 13, 61 and 117 from tumor DNA using the polymerase chain reaction technique (PCR). The amplification products were subjected to direct DNA sequence analysis to identify point mutation within these activating codons. The results showed that tumors induced with the nongenotoxic hepatocarcinogens phenobarbital, chloroform or ciprofibrate have a much lower frequency of H-ras gene activation than do those that arose spontaneously or those induced with the genotoxic carcinogen benzidine-2 HCl. 127 MATERIALS AND METHODS Tumor Induction. Spontaneous liver tumors were obtained from control male B6C3F1 mice collected at the terminal necropsy of 2-year bioassays for carcinogenicity conducted at the Environmental and Toxicology Research Laboratory of The Dow Chemical Company (Midland, MI). Chemically induced liver tumors were obtained from independently treated groups of male B6C3F1 mice administered either benzidine-2 HQ (120 ppm, drinking H20, 1 year), phenobarbital (0.05%, drinking H20, 1 yr), chloroform (200 mg/ kg corn oil gavage, 2X weekly for 1 yr) or ciprofibrate (0.0125% diet, 2 yr). Animals were put on test at 6 weeks of age and necropsied after 18-24 months. At necropsy tumors were excised and a section of each tumor was fixed in neutral phosphate buffered 10% formalin. Tissues were processed by conventional techniques, embedded in paraffin, cut at approximately 6pm, stained with hematoxylin and eosin and examined using a light microsc0pe. The remainder of the tumor was frozen on dry ice and stored at -70°C until DNA was isolated. Oligonucleotide Primer Preparation. Oligonucleotide primers used in PCR amplification and DNA sequencing were synthesized using the phosphoramidite method on an Applied Biosystems Model 380A DNA synthesizer. Detritylated primers were purified by HPLC using a 101i RP-300 reverse phase column (Brownlee Laboratories). Elution of primers from the column was achieved with a linear 7-17% acetylnitrile gradient (50 min., 1 128 ml/ min.) in 0.1 M tetramethylammonium acetate (pH 7.0). HPLC fractions containing the purified primer were pooled, evaporated to dryness, reconstituted with distilled H20 and purified by ultrafiltration using a Centricon-3 (Amicon, W. R. Grace, Danvers, Ma.). PCR Amplification. Total genomic DNA was isolated from liver tumor tissue as previously described (5). Prior to PCR amplification approximately 2 ug of tumor DNA was digested with 2 units of Bam HI restriction endonuclease (37°C, 30 min) as recommended by the manufacturer (Bethesda Research Laboratories). Restriction endonuclease digestion of the genomic DNA resulted in an increase amplification yield of the H-ras gene sequences presumably by increasing the template accessibility to the PCR polymerizing enzyme (unpublished studies, T. Fox). PCR amplification of the digested tumor DNA was conducted in a total volume of 50p] containing 50 mM KCl, 10 mM Tris-HCl (pH 8.3), 15 mM MgC12, 400 M dATP, dT'I'P, dGTP, dCTP, approximately 1 ug each of 3' and 5' oligonucletotide primers and 2.5 units of Taq polymerase (Perkin-Elmer Cetus)(10). The reaction mixture was incubated for 5 min at 94°C to ensure complete denaturation of genomic DNA. Primers were than allowed to hybridize to the template DNA at 65°C for 1 min. Amplification was initiated with the addition of Taq enzyme followed by 30 cycles of successive incubations for 1 min each at 65°C (annealing) 70°C (polymerization) and 94°C (denaturing). Confirmation of amplification was assessed by agarose gel 129 electrophoresis. 10111 of the PCR reaction mixture was loaded onto a 3% Nusieve (FMC) 1% agarose gel and DNA bands were visualized by ethidium bromide staining. Sequencing. The protocol for DNA sequence analysis was adopted from that of Yang e_t s1. (11). Following the termination of amplification the PCR reaction mixture was purified by centrifugation through a Centricon-30 to remove unincorporated nucleotides and excess oligonucleotide PCR primers. The Centricon-30 retentate was sequenced directly using a modification of the Sanger dideoxy chain termination protocol (12 ). Five microliters of the PCR amplified sample was added to 15111 of a solution containing 40 mM Tris-HCl (pH 7.5), 20 mM MgC12, 50 mM NaCl and approximately 5 pmoles of 32P end- labeled sequencing primer. Sequencing primers were end-labeled with (7-32P) ATP by incubation in 60 mM Tris-HCl (pH 7.8) 10 mM MgC12, 330 mM ATP, 15 mM 2- mercaptoethanol and 15 units of polynucleotide kinase at 37°C for 30 min. The reaction mixture was centrifuged through a N u-Clean D25 disposable spin column aspdescribed by the manufacturer (IBI). Specific activities of the sequencing primers routinely ranged from 1-5 X 108 CPM/ ug. The mixture containing the amplified DNA and sequencing primer was heated for 5 min. at 96°C, cooled on ice, and 2 units of sequenase (United States Biochemical) was added. The sequencing reaction was initiated by 130 adding 3.5 ul of this mixture to 1 of 4 tubes containing 2.5 ul of the ' appropriate combination of deoxynucleotides/ dideoxynucleotide at a 10:1 ratio. Deoxynucleotide concentrations were 33 Md and dideoxynucleotides were 3.3 M. The tubes were incubated at 37°C for 15 min. and the reaction was terminated by the addition of 4 pl of 95% formamide containing 0.05% bromophenol blue and xylene cyanol FF tracking dyes. The reaction tubes were heated at 96°C for 3 min, cooled on ice and 2.5 ul was loaded onto a 8% polyacrylamide/BM urea gel and electrophoresed at 40 m Amps. The gel was transferred to Whatrnan 3 mm paper, dried under vacuum and exposed to Kodak XAR-S film at -70°C. for 24-48 hrs. RESULTS Tumor Induction. Tissue used for analysis in this study was derived from male B6C3F1 mice liver tumors occurring spontaneously or induced by treatment with the genotoxic hepatocarcinogen benzidine-2 HCl or the nongenotoxic hepatocarcinogens phenobarbital, chloroform or ciprofibrate. Mice bearing spontaneous tumors, obtained from the control groups of 6 different bioassays, had an average tumor incidence of 42% with 22% of the tumor bearing animals having multiple liver tumors (Table 1). A treatment related effect was observed for all compounds within the treatment groups as indicated by the increase in percent tumor incidence and tumor multiplicity compared to the control animals (Table ’1). Tumors ranged in size from approximately 1 mm to 2.5 cm in diameter. 131 H-ras Gene Activation. The activation of the H-ras gene in a variety of human and rodent tumors has been shown to involve point mutations at codons 12, 13, 61, or 117 (10,14). The extent of H-ras gene activation in the B6C3F1 mouse liver tumors of the various treatment groups was assessed by determining the presence or absence of point mutations within these activating codons. Codons 12 and 13, 61 and 117 reside in exons 1, 2 and 3, respectively, of the H-ras gene (15). Point mutations in these codons were identified by direct sequence analysis of amplified regions within the corresponding exons from genomic tumor DNA. These exon regions were amplified using the polymerase chain reaction (PCR) technique. The amplified regions of the H-ras gene with the corresponding PCR and sequencing primers are illustrated in Figure 1. Sequence analysis of the H-ras gene was carried out on a total of 151 tumors from the combined groups. Mutations were detected only in codon 61 and not at codons 12, 13 or 117 in any of the tumors. As a result the percentage of tumors with a mutation in codon 61 is equivalent to the percentage of tumors with an activated H-ras gene. As can be seen in Table 2, the percentage of tumors with an activated H-ras gene in the spontaneous and benzidine-2 HCl-induced tumors is considerably higher than in the tumors induced by treatment with phenobarbital, chloroform or ciprofibrate. Analysis of spontaneous tumors revealed that 64% contained a point mutation in codon 61. A similar response was observed in the benzidine-2 HCl-induced tumors, where 59% 132 tumors had codon 61 mutations. This is in contrast to the tumors induced by the nongenotoxic agents where activating H-ras gene mutations were detected at a much lower frequency, i.e., phenobarbital 7% , chloroform 21 % and ciprofibrate 21 %.. Histopathological examinations were conducted on all tumors except those that were too small to forfeit tissue for analysis. A correlation between the incidence of H-ras gene activation and the development of either a hepatocellular adenoma or hepatocellular carcinoma was made and the data summarized in Table 3. There was no statistically significant difference between the frequency of H-ras gene activation in the hepatocellular adenomas and carcinomas (statistical analysis determined using a chi square contingency table test). Histopathological examination of the spontaneous tumors and the tumors induced with benzidine-ZHCI, phenobarbital, and chloroform did not reveal any significant differences in morphology or staining characteristics. The ciprofibrate induced tumors were more eosinophilic as were the surrounding normal hepatocytes. This observation would be expected from a treatment related increase in cytoplasmic peroxisomes. The mutational spectrum of activating codon 61 mutations within the H-ras gene was determined and the data are summarized in Table 4. The normal nucleotide sequence for codon 61 in the mouse H-ras gene is CAA. The activating mutations observed in all tumor groups involved a single base pair substitution changing this sequence to AAA, CGA or CTA. 133 Representative examples of sequence gels exhibiting these three mutations are shown in Figure 2. The most frequently detected mutation was a C-G to AT transversion at the lst nucleotide position. The frequency of this base pair substitution in the spontaneous tumors and in the tumors induced with benzidine 2-HCl, phenobarbital, chloroform and ciprofibrate was 59%, 85%, 100%, 80% and 88% respectively. An A-T to CC transition or a A-T to T-A transversion at the second nucleotide position occurred in these same groups at a frequency of 34%, 8%, 0%, 0%, 12%, and 6%, 8%, 0%, 20%, 0% respectively. Many of the animals in the chemically induced tumor groups had multiple tumors as is indicated in Table 1. A comparison of the H-ras codon 61 nucleotide sequence of the individual tumors from animals with multiple tumors revealed that in approximately one-half of those animals examined a different codon 61 sequence was observed. DISCUSSION Mutational analysis of tumor DNA from spontaneous tumors and tumors induced with benzidine-2 HCl, phenobarbital, chloroform, or ciprofibrate revealed activating mutations exclusively in codon 61. Only 3 activating point mutations within this codon were detected; a CG to A-T transversion at the lst nucleotide position, an A-T to G-C transition or an A-T to T-A transversion at the second nucleotide position. These point mutations 134 resulted in the amino acid substitution of glycine in the H-ras P21 protein with lysine, arginine or leucine respectively. The predominant mutation in all treatment groups was the point mutation converting the normal codon 61 sequence of CAA to AAA. It has been demonstrated by Strauss et a1 (17) that during DNA synthesis on a noninstructional template the most frequent polymerase error involves the preferential insertion of an adenine, the"A rule". In the spontaneous tumors the high frequency of CC to AT transversions observed at this nucleotide position may be attributed to this type of polymerase error. The frequency of AAA codon 61 mutations in DNA from tumors taken from benzidine2 HCl treated animals was slightly higher than that observed in the DNA from spontaneous tumors. Chi square analysis showed that this increase was not statistically significant. Some of the tumors carrying this Specific mutation could be of spontaneous origin considering the high frequency of this mutation observed in spontaneous tumors. However, this increase could be due to the mutagenic activity of benzidine- 2HC1. Benzidine ~2HCl is a potent genotoxic agent which has been shown to cause primarily C8 adducts on guanine (18). The formation of such an adduct on the guanine opposite cytosine at the first position in codon 61 could be expected to cause a CG to AT transversion. This transversion could occur either by the preferential insertion of an adenine across from a noninstructive base or as a result of the adduct allowing stable base pairing between the modified guanine with an adenine as was reported recently for guanine carrying an AP residue at the C8 position (19). 135 Several investigators have also determined the mutational spectra of codon 61 within the H-ras gene from B6C3F1 mouse liver tumors induced with various genotoxic carcinogens. Wiseman et al. (9) induced B6C3F1 liver tumors with N-hydroxy-Z-acetylaminofluorene, vinyl carbamate, or 1'- hydroxy-2',3', dehydroestragole. These agents produce codon 61 mutational spectra that are distinct from the mutational spectra observed in spontaneous tumors. A similar analysis by Reynolds et al. (8) of tumors induced with furan and furfural observed mutations not only in codon 61 but at codons 13 and 117, two activating codons that have not been detected previously in tumor DNA. The carcinogenic agents used in these studies therefore produce unique chemical mutation spectra within the H-ras gene which is consistent with their apparent genotoxic activity. It has been suggested that the evaluation of the mutational spectra within the H-ras gene of B6C3F1 mouse liver tumors may be useful in determining the in viva genotoxic potential of an agent which is positive for tumorigenicity in the rodent bioassay (10). This type of analysis may prove valuable in selected circumstances but there are limitations that must be realized; This is illustrated by the analysis of liver tumors generated in the B6C3F1 mouse with the potent mutagenic carcinogen N-diethylnitrosamine. The codon 61 mutation spectrum observed in these tumors was similar to the mutation spectrum observed in spontaneous tumors (20). Consequently, the presence or absence of genotoxic activity for this particular chemical cannot be concluded using this procedure. This type of analysis is useful only if the mutational spectra of the chemically induced tumors deviates significantly from the spontaneous tumor spectrum. 136 Because the spontaneous tumor spectrum is used as a benchmark for this analysis, a high degree of confidence in this spectrum is an obvious prerequisite. One of the objectives of the present study was to examine a large number of spontaneous tumors to increase the statistical confidence in this spectrum. Our results have indicated that the mutational spectrum observed in the spontaneous liver tumors is similar to the mutational spectrum previously reported for 15 spontaneous tumors (10). The inclusion of this data to the preexisting data base should increase the degree of confidence in this spectrum. Analysis of the codon 61 mutational spectra of the tumors induced with the nongenotoxic agents used in this study was not meaningful because so few of the tumors contained codon 61 mutations. However, it was noted that the most frequently observed mutation in these tumors, a CC to A-T base transition at the first nucleotide position, was the most common mutation detected in the spontaneous tumors. Analysis in the frequency of H-ras gene activation did reveal significant differences between the tumors induced with the nongenotoxic carcinogens compared to the spontaneous tumors or tumors induced with the genotoxic agent benzidine - 2HCL. The H-ras gene was activated to a similar frequency in spontaneous and benzidine -2 HCl induced tumors (64% and 59% respectively). The frequency of H-ras gene activation was significantly lower in tumors induced with phenobarbital, chloroform, and ciprofibrate (7%, 21 %, 21% respectively). If one determines the number of tumors within the nongenotoxic induced groups that would be of spontaneous origin (42%), and then calculates 137 the subgroup of these that can be expected to exhibit an H-ras activation, i.e. 27% (64% of 42%), this value approximates the number of tumors observed within these groups having an activated H-ras gene. In other words, the number of tumors in these groups that contained an activated H-ras gene is the number that would be expected if those animals had received no treatment. This suggests that the tumors induced with the nongenotoxic agents having an activated H-ras gene are probably of spontaneous origin and, therefore, the increase in tumor frequency above background caused by exposure to these nongenotoxic agents results from their interaction with a population of non-H-ras activated cells. We concluded that these nongenotoxic agents do not cause mutations within the H- ras gene, and that they increase the frequency of mouse liver tumor development by a mechanism that differs from the process occurring spontaneously or after treatment with a genotoxic carcinogen. Activation of this gene was detected in hepatocellular adenomas as small as 1 mm in diameter. Balmain et al (21) have detected an activated H-ras gene in chemically induced benign mouse skin papillomas. The detection of an activated H-ras gene in very small hepatocellular adenomas as well as in benign skin papillomas suggests that activation of this gene may be an early initiating event in tumorigenesis. The high frequency of H-ras gene activation in spontaneous tumors also indicates that this gene plays an important role in B6C3F1 mouse liver tumor development. However, in approximately 35% to 40% of the spontaneous and benzidine.2HCl induced tumors and in approximately 80% to 90% of the tumors induced with the nongenotoxic agents did not have a detectable activated H-ras gene. This indicates that alternative 138 pathways independent of H-ras gene activation can also lead to hepatic tumor development. The data suggest that there are at least two populations of cells which can give rise to tumors within the liver, one consisting of cells having an activated H-ras oncogene and another which does not. The data do not preclude the possibility that cells giving rise to tumors that do not have an activated an H- ras gene may have other activated ras genes. (i.e. K-ras, N-ras). Reynolds st 31., did detect an activated K-ras gene in 7% and 15% of liver tumors in the B6C3F1 mouse induced with furfural and furan respectively (10). The three nongenotoxic agents used in this study have diverse pharmacologic and toxicologic activities. Phenobarbital has general hepato- stimulatory properties resulting in increased gene expression with an associated induction of cellular proliferation (22). Animal tumors induced with chloroform are accompanied by cell necrosis, death, and subsequent regeneration (23). Although cellular proliferation alone is probably not sufficient to cause tumors it has been shown to be a necessary prerequisite for the fixation of the initiating event in cell culture and it is believed to be important for the clonal expansion of the initiated cell population in viva (24, 25). Peroxisome proliferators, such as ciprofibrate, have been hypothesized to be responsible for tumor induction by various mechanisms. One of these suggests that they induce indirect secondary damage to DNA as a consequence of an increased production of oxygen free radicals, however, ciprofibrate has not been shown to cause mutations in vitra (26). Alternatively it has been proposed that such agents stimulate the growth of preneoplastic foci suggesting that they are acting as tumor promoters (27). 139 Even though the biological activity of these nongenotoxic animal tumorigens is diverse, their tumorigenic activity is only observed at doses that cause perturbations in the cells' normal biochemistry and/ or physiology. This implies that at lower doses they will not cause tumors. In summary, results from the present study indicate that the nongenotoxic agents phenobarbital, chloroform and ciprofibrate induce liver tumors in the B 6C3F1 mouse by a process that does not require the activation of the H-ras gene. This data suggests that at least two different populations of cells, one requiring H-ras activation and the other independent of this event, can give rise to liver tumors in these animals. The nongenotoxic agents used in this study apparently increase B6C3F1 mouse liver tumors by interacting with a population of cells that do not contain an activated H-ras gene. Additional studies comparing tumors with and without an H-ras gene mutation will be required to elucidate the molecular and cellular differences between these cell populations. Understanding these differences should provide insight into how various classes of carcinogens interact with cells to give rise to liver tumors in the B6C3F1 mouse. Such results will be useful in extrapolating bioassay results using this strain of mouse to evaluate potential chemical carcinogenic risk for humans. 140 ACKNOWLEDGEMENTS The authors would like to thank our colleague Dr. Jia-ling Yang, presently at the Institute of Biomedical Sciences, Academia Sinica, Taipei, Taiwan for here advice on the DNA sequence analysis. We also thank Dr. Peggy Schott for her technical advice concerning the DNA oligonucleotide synthesis, Sherry Pagels for technical assistance and Jean Shuler for typing the manuscript. 141 REFERENCES Chen, C., NTP technical report on the toxicology and characterization studies of dimethylmorphoenophosphoramidate in F344 / N rats and B5C3F1 mice. Cas No. 597-25-1, 1984. Ward, J. M., Goodman, D. G., Squire, R A., Chu, K. C., and Linhart, M.S., (1979), Neoplastic and nonneoplastic lesions in aging (57BL/ 6N x C3H/HeN) F1 (B6C3F1), J. Natl. Cancer Inst, 63(3): 849-854. Ashby, J. and Tennant, R. W. Chemical structure, Salmonella mutagenicity and extent of carcinogenicity as indicators of genotoxic carcinogenesis among 222 chemicals tested in rodents by the US. NCI/NTP. Mutation Bis, 204:17-115, 1988. Cellufilr and Molecul_ar Biology of Cancer, Editors Franks, L. M., and Feich, N., Oxford University Press, New York, NY, 1986. Schulte-Hermann, R, Timmermann-Trosiemer, I., Schuppler, 1., (1983), Promotion of spontaneous preneoplastic cells in rat liver as a possible explanation of tumor production by nonmutagenic compounds, Cancer Res, 43: 839-844. 10. 142 Frenkel, K., and Chrzan, K., (1987), DNA base modification by tumor promoting activated polymorphonuclear leukocytes, Carcinogenesis 8: 455-460. Fox, T. R, Schumann, A. M., Watanabe, P. G. Activation of a cellular proto-oncogene in spontaneous liver tumor tissue of the B6C3F1 mouse. Arch Toxicol. Suppl., 10:217-227, 1987. Reynolds, 5. H., Stowers, S. 1., Maronpot, R R, Anderson, M. W. and Aaronson, S. A. Detection and identification of activated oncogenes in spontaneously occurring benign and malignant hepatocellular tumors of the 86C3F1 mouse. Proc. Natl. Acad. _S£i_. USA, 86:33-37, 1986. Wiseman, R W., Stowers, S. 1., Miller, E. C., Anderson, M. W. and Miller, 1. A. Activating mutations of the C-Ha-ras proto-oncogene in chemically induced hepatomas of the male B6C3F1 mouse. PLog. Na_tl. _A_csc_l. S_c_i. USA, 83:5825-5829, 1986. Reynolds, 5. H., Stowers, 5.1., Patterson, R M., Maronpot, R R, Aaronson, S. A. and Anderson, M. W. Activated oncogenes in B6C3F1 mouse liver tumors: implications for risk assessment. Science, 237:1309- 1316, 1987. 11. 12. 13. 14. 15. 16. 143 Mullis, K. and Faloona, F. Specific synthesis of DNA in vitra via a polymerase catalyzed chain reaction. Methods Enzmol., 155:335-350, 1987. Yang, 1-L., Maher, V. M., McCormick, J. 1. Amplification and direct mucleotide sequencing of cDN A from the lysate of low numbers of diploid human cells. Gene (in press). Sanger, F., Nicklen, S. and Coulson, A. R DNA sequencing with chain terminating inhibitors. Proc. Natl. Acad. E. USA, 74:5463-5467, 1977. Barbacid, M. Ras genes, Annual Rev. Biochem., 56:779-827, 1987. Fasano, O., Taparowsky, E., Fiddes, 1., Wigler, M. and Goldfarb, M. Sequence and structure of the coding region of the human H-ras-l gene from T24 bladder carcinoma cells. _1. Molecular and Applied §e_nt. 2:173- 180, 1983. Reddy, E. P., Lysman, D., Andersen, P. R, Tronick, S. R. and Aamson, S. A. Nucleotide sequence analysis of the Balb/c murine sarcoma virus transforming gene. 1. Virology, 53:984-987, 1985. 17. 18. 19. 20. 21. 144 Strauss, B., Rabkin, F., Sagher, D., Moore, P. The role of DNA polymerase in base substitution mutagenesis on noninstructional templates. Biochimie 674:829-838, 1982. Josephy, P. D., Benzidine: mechanism of oxidative activation and mutagenesis. m. Proc., 45(10):2465-2470, 1986. Norman, D., Abuaf, P., Hingerty, B.F., Lide, D., Grunberger, D. ,Broyde. S. and Patel, D. J. N MR and Conformational characterization of the N- (Deoxyguanosine 8-yl)aminofluorine adduct-[(AF)G] opposite adenosine in DNA; (AFG[syn]- A[anti] pair formation and its' pH dependence. Biochemist_ry, 28: 1989 (in press) Stowers, S. J., Wiseman, RW., Ward, 1.M., Miller, E.C., Miller, I.A., Anderson, M.W., Eva, A., (1988), Dection of activated proto-oncogenes in N-nitrosodiethnylamine induced liver tumors: a comparison between B6C3F1 mice and Fischer 344 rats, Carcinogenesis, 9 (2): 271-276. Balmain, A., Ramsden, M., Bowden, G. T, and Smith, 1. Activation of the mouse cellular Harvey-ras gene in chemically induced benign skin papillomas. Nature 307:658-660, 1984. Peraino, C., Fay, R. J. M., Staffeldt, E. and Christopher, J. P. Comparative enhancing effects of phenobarbital, amobarbital, diphenylhydantoin and 24. 26. 27. 145 dichlorodiphenyl-trichloroethyane on 2-acetylaminofluorene induced hepatic tumorigenesis in the rat. Cancer Research, 35:2884-2890 (1975). Reitz, R H., Fox, T. R and Quast, J. F. Mechanistic considerations for carcinogenic risk estimation: chloroform. Envirn. Health Perspectives, 46:163-168 (1982). Kakunaga, T. The role of cell division in the malignant transformation of mouse cells treated with 3-methylcholanthrene. Cancer Research 35:1637-1642 (1975). Mironescu, 5., Love, R. DNA synthesis and transformation induced in density-inhibited cultures of hamster embryo cells by the carcinogen benzo(a)pyrene. Cancer Research 34:2562-2570 (1974). Reddy, 1. K and Lalwani, N. D. Carcinogenesis by hepatic peroxisome proliferators: Evaluation of the risk of hypolipidemic drugs and industrial plasticizers to humans. CRC Crit. Rev. Toxicol. 12:1-58 (1983). Cattley, R. C. and Popp, J. A. Differences between the promoting activities of the peroxisome proliferator WY-14,643 and phenobarbital in rat liver. Cancer Research, 49:3246-3251 (1989) 146 Figure Legend Figure 1. . Nucleotide sequence of the coding strand for the H-ras gene Figure 2. which was PCR amplified and subsequently sequenced. Fragment I represents the region of exon 1 containing codons 12 and 13. Fragment II is region of exon 2 containing codon 61. Fragment III is the entire sequence of exon 3 containing codon 117. The darkly shaded areas indicate position of the 5' and 3' PCR primers. Lightly shaded areas represent the sequencing primers. Representative DNA sequence of the H-ras gene codon 61 region derived from PCR amplified liver tumor DNA. (a) Normal codon 61 sequence (CAA) derived from a ciprofibrate induced tumor. (b) CG —> A-T transversion mutation at codon 61 (AAA) from ciprofibrate induced liver tumor. ‘ (c) A-T _> G-C transition (CGA) from benzidine-ZHCl induced liver tumor. (d) A-T —> T-A transversion (CTA) from spontaneous liver tumor. 147 EmEca 0:65:08 3 295a mom . .. m8_> mEoocmw 2:56 928 E0: 32.8 mm .. I I I a a c a i . n t e t. o a e a s . es. 148 Figure 2. ng 149 Table 1: Incidence of spontaneous and chemcially induced liver tumors in male B5C3F1 mice. % of tumor bearing % of animals Treatment Animals animals with used to induce Number of with with . multiple tumors animals tumorsa tumors tumors None 293 124 42 22 (Spontaneous) Benzidine - 27 23 85 39 2HC1 Phenobarbital 25 17 68 65 Chloroform ‘ 17 15 80 80 Ciprofibrate 31 29 94 79 a) All animals were necropsied after 18-24 months of age and liver tumors were removed for further analysis. 150 Table 2: Frequency of activation of the H-ras gene at codon 61 in male B5C3Fl mouse liver tumors. Treatment used to Number of tumors Tumors with an induce tumors analyzed“ activaed H-rasa None 50 32 (64%) (Spontaneous) Benzidine ~2HCl 22 ' 13 (59%) Phenobarbital 15 1 (7%) Chloroform 24 5 (21%) Ciprofibrate 39 8 (21%) a) Total genomic DNA was isolated from liver tumor tissue. Defined regions of exon 1, 2, and 3 of the H-ras gene incorporating codons 12, 13, 61, and 117 were PCR amplified and sequenced directly. The sequence was determined for 97 of the total 190 codons of the H-ras gene. 151 Table 3: Frequency of H-ras gene activation in hepatocellular adenomas and carcinomas in male B5C3F1 mice. % of Adenomas with % of Carcinomas with Treatment used to activated activated induce tumors H-rasa H-rasa None (Spontaneous) 61% (11/18) 64% (9/ 14) Benzidine -2HCl 53% (8/ 15) 67% (4/6) Phenobarbital 1 1 % (1 / 9) 0% (0/ 5) Chloroform 15% (2/ 13) 33% (3/ 9) Ciprofibrate 21% (6/ 29) 22% (2/ 9) a) A portion of each tumor was excised, fixed at necropsy in neutral phosphate buffered 10% formalin, processed by standard techniques, and embedded in prarfin. Sections were cut (6 pm) , stained, and evaluated by light microscopy. 152 Table 4: Mutation spectrum in codon 61 (CAA) of the H-ras gene in male B5C3F1 mouse liver tumors. Number of tumors with specific Treatment used . Tumors with an codon 61 mutations (%)a to induce activated H-ras tumors gene AAA CGA CTA None 32 19 (59) 11 (34) 2 (6) (spontaneous) Benzidine . 2HCl 13 11 (85) 1 (8) 1 (8) Phenobarbital 1 1 (100) 0 0 Chloroform 5 4 (80) 0 l (20) Ciprofibrate 8 7 (88) 1 (12) 0 a) Codon 61 sequence was determined as described in Table 2. 153 CHAPTERHLAMQDELEQW nu rant! 0' tie 01L .t - MNT Determining the mechanisms involved in the development of B6C3F1 mouse liver tumors is not only important for interpreting cancer bioassay data, but may also provide a useful model system for understanding carcinogenesis in general. Even though the specific events of neoplastic development in the mouse liver have not been completely elucidated, results from a number of recent studies have provided important insights into the process of spontaneous and chemically induced hepatocarcinogenesis in this strain of mouse. The B6C3F1 mouse is a hybrid obtained through the genetic cross between a male C3H/He1 and a female C57BL/61. These two parental strains have widely differing susceptibilities to the development of both spontaneous and chemically induced liver tumors. Within 15 months of age nearly 50% of the C3H/HeJ mice will develop spontaneous liver tumors (Andervant, 1950). In contrast less that 5% of the C57BL/ 61 mice will have liver tumors by 2 years of age (Grasso and Hardy, 1975). This difference in spontaneous tumor susceptibility is also exhibited in the chemical induction of liver tumors. A 15 fold greater multiplicity of tumor formation is observed in the C3H/ He] mouse compared to the C57BL/ 6] mouse after treatment with 4- aminoazobenzene or N,N-dimethyl-4-aminoazobenzene (Declos et al., 1984). The C3H/HeJ mouse is also 20-50 fold more susceptible than the C57BL/6J 154 mouse to the induction of liver tumors by N-ethyl-N-nitrosourea (ENU) or N,N-diethylnitrosamine (DEN) (Drinkwater and Ginsler, 1986). The susceptibility of the C3H/ He] mouse to develop both spontaneous and chemically induced liver tumors is presumably due to a genetically transmitted predisposition gene(s). A study to delineate the genetic basis for this susceptibility was recently reported by Drinkwater (Drinkwater and Ginsler, 1986). DEN induced tumor multiplicity was used as a genetic marker in segregating cross analysis between various susceptible and resistant strains. In these studies it was determined that tumor susceptibility was due to allelic differences for at least 2 independent loci (Drinkwater and Ginsler, 1986). A single locus, however, appears to be responsible for approximately 85% of this sensitivity. This locus has been denoted the Hcs gene (Hepato-carcinogen sensitivity). The dominant and recessive character of this locus was evaluated by exposing a series of recombinant inbred mice strains to ENU and determining the relative susceptibilities to develop liver tumors. Three different susceptibility phenotypes were observed. One was very sensitive, one was resistant and one was intermediate to the others. This data indicates that the alleles of the Hcs locus in the C57BL/ 61 and C3H/HeJ mice are semi- dominant. The susceptibility of the B6C3F1 mouse to develop spontaneous and chemically induced tumors falls between the two parents (Becker, 1982, Drinkwater and Ginsler, 1986). The intermediate tumor sensitivity of this mouse is a logical outcome given the semi-dominant nature of the Has gene. 155 At what stage in the carcinogenic process does this gene exert its influence? To address this issue, Drinkwater and his collaborators conducted several studies ‘to determine whether the H as locus affects initiation or post- initiation events in the liver. Newborn C5le 6] and C3H/ He] mice were injected with [C 14] DEN and the relative levels of DNA alkylation was determined. The results demonstrated that there was no significant difference in the extent of ethylation of hepatic DNA, or in the relative levels of N-7-ethylguanine or Oé-ethylguanine adducts between the two strains (Drinkwater and Ginsler, 1986). The genetic influence of this locus does not, therefore, appear to be at the level of DNA modification by hepatic carcinogens. This suggests that the Has gene may be acting at the promotion stage of liver tumor development. To investigate this possibility, the comparative development of putative preneoplastic lesions induced with ENU was made between each strain (Hanigan et al., 1988). In both strains an increase in the number and size of these lesions was observed. At 12 weeks of age there was approximately a 25 fold greater number of foci in the C3H/ He] mice compared to the C57BL/6J mice . This data could be interpreted to mean that the Hcs gene is influencing the initiation stage of tumor development. The difference however had decreased to only 5 fold by 20 weeks of age. On the other hand, the ratio in the mean diameter of these lesions between the C3H/ He] and C57BL/6J mice continued to increase with time, indicating that the foci in the C3H/ He] mice were growing faster than the foci in the C57BL/6J mice. This rate of growth was determined to be about 70% greater 156 in the C3H/ He] mice compared to the C57BL/6J mice. Taking into consideration both the divergence over time in the size of the foci with the convergence in the total number of foci, the data suggests that the Hcs gene is probably acting at the level of promotion. From these data it was concluded that the greater susceptibility of the C3H/ He] and B6C3F1 mice to develop liver tumors compared to the C57Bl/6] mouse is the result of a strain dependent difference in the rate of grth in preneoplastic lesions caused in part by the Has gene. . It is now evident that neoplastic development within the liver of the B6C3F1 mouse involves the participation of more than the Hcs gene. As indicated in the discussion from previous sections of this thesis, the H-ras oncogene is found in a majority of spontaneous tumors and tumors induced by exposure to various genotoxic carcinogens. This oncogene appears to play an important role in the development of these tumors and most likely at the initiation stage of tumor development. It does not however appear to be essential for hepatocarcinogenesis because not all of these tumors contain an activated H-ras gene. This is even more evident in tumors induced with certain nongenotoxic carcinogens where the vast majority do not contain this activated oncogene. The transforming function exerted by the H-ras gene apparently can be functionally substituted for by the activation of some other gene(s). Given this information, what type of model can be constructed to explain the sensitivity of the B6C3F1 mouse to develop liver tumors after exposure to various nongenotoxic agents? Many nongenotoxic carcinogens 157 exhibit their tumorigenic action only at doses which are high enough to induce cellular toxicity. As was discussed in the section on nongenotoxic carcinogens, an enhanced level of cell division can occur as a consequence of a toxicity induced degenerative / regenerative response in the target organ. During periods of rapid cellular proliferation additional mutations in the DNA can occur because less time is available for premutagenic lesions to be repaired (Maher et al., 1979). Consequently a greater number of these premutagenic lesions will become fixed as mutations. This higher incidence of spontaneous mutations may create a greater number of initiated cells. With an expanded population of initiated cells the probability increases that one of these cells will give rise to a tumor. Alternatively, if these mutations occur in previously initiated cells then they may be advanced to a higher state of neoplastic transformation and therefore more prone to develop into a tumor. Alternatively, the nongenotoxic carcinogens may be exerting their tumorigenic action by acting as hepatic tumor promoters. This promotional activity may be exerted in several different ways. It has been suggested that these agents may provide a promotional stimulus by increasing hepatic cellular proliferation for example through a degenerative/ regenerative hyperplasia (Schulte-Hermann et al., 1982; Mirsalis et al., 1985). An increase in the level of liver cell proliferation could result in an expansion of the previously initiated cell population. This expansion in the number of initiated cells increases the probability that one of these cells may progress further towards neoplastic development and eventually give rise to a tumor. 158 Some nongenotoxic agents can stimulate cell division by mechanisms other than by inducing cellular toxicity. Many of the classic tumor promoting phorbol esters have have inherent mitogenic activity that can stimulate cell proliferation. These agents have been demonstrated to be potent activators of protein kinase C (Couturier et al., 1984). The activation of this enzyme can have many physiological consequences including the induction of cell division. Some nongenotoxic agents may also exert their promotional activity by altering the expression of gene transcription within the initiated cell. For example, the polychlorinated biphenyl 2,3,7,8-tetrachlorodibenzodioxin (TCDD), along with other halogenated aromatic hydrocarbons, has been shown to be a potent liver tumor promoter in rats (Pitot et al., 1980; Jenson et al., 1982). These materials bind to high affinity cyt0plasmic receptors which then can interact with specific promoter regions within the DNA causing alterations in gene transcription. The consequences of this altered gene activity on neoplastic transformation are not well understood but changes in gene expression are considered to be an important aspect of tumor promotion. Recent studies by Drinkwater et al. (in press) examined the relationship between the promoting activity in one of these polyhalogenated biphenyl tumor promoters and the Has gene. They compared the susceptibility to develop tumors in ENU initiated C57BL/ 61 and C3H/ He] mice with the promotional activity of the TCDD congener 3,4,5,3',4',5'-hexabromobiphenyl (HBB). Their results demonstrated that the Hcs gene can act synergistically 159 and independently with this tumor promoter to enhance the induction of liver tumors in the sensitive C3H/ He] mouse. The increase in B6C3F1 mouse liver tumors as a result of chronic exposure to nongenotoxic agents may be the result of an increase in the number of initiated cells and/ or the promotion of existing spontaneously initiated cells. Both of these actions could be the consequence of an increase in cellular proliferation or through alterations in gene expression. These activities when coupled to the synergistic action of the Has gene have the apparent enhancing effect of increasing the tumor response in these animals. One perplexing question still remains to be addressed. Why do the tumors induced with these agents have a much lower frequency of the activated H-ras gene compared to spontaneous tumors or tumors induced with genotoxic carcinogens? This is a difficult question to answer. It was originally hypothesized that this class of chemicals caused an increase in liver tumors by simply enhancing the spontaneous process of tumor development. As a result treatment by these agents would increase the tumor incidence, but the distribution of tumors with or without an activated H-ras gene would be the same as is in the control animals. The few tumors observed within the nongenotoxic treatment groups that contain an activated H-ras gene are approximately the number that would be expected if those animals had received no treatment. This suggests that the increase in tumor frequency was due to an enhancement in the tumorigenic conversion of those cells which did not contain an activated H- ras gene. This implies that the increase in liver tumors within the 160 nongenotoxic treatment groups was the result of a preferential stimulation of the non H-ras containing cells by these agents. How might this occur? To address this question we must consider the process of initiation. It is not known whether there is more than one type of initiating. event. The present analysis supports the concept that there may be at least two types of initiation within the liver; activation of the H-ras gene and another not involving the activation of the H-ras gene. It is also not known whether these different initiation events are of equal potency. One type of initiated cell may have a greater potential to progress towards neoplastic transformation that another. Because mare of the spontaneous tumors contain an activated H-ras gene than do not, it could be argued that the activation of the H-ras gene is a more potent initiation event (assuming all initiating events occur with equal frequency). The cells containing an activated H-ras gene exist at a more advanced state of neoplastic progression than do non-H-ras initiated cells. Consequently they require fewer additional changes to arrive at a state of complete neoplastic conversion. This explains why a higher percentage of spontaneous tumors contain an activated H-ras gene. Suppose that (1) the nongenotoxic agents influence tumor development only at a certain step in the tumorigenic process and (2) only the cells containing an activated H-ras have progressed beyond the point of influence by the nongenotoxic agents. If this is true, then the progression of the H-ras containing cells to a more advanced level of neoplastic transformation is not influenced by the exposure to these agents. 161 On the other hand, the ne0plastic progression of those cells without an activated H-ras gene is influenced by exposure to these chemicals. Perhaps exposure to these agents alters the expression of specific genes which elevates those cells to a level of neoplastic conversion that is comparable to the H-ras containing cells. Once they have reached that point both cell types now have an equal probability of further neoplastic progression. However, the exposure to these agents has now increased the number of cells containing an activated H-ras gene to this elevated level of neoplastic transformation. This increases the probability that one of these cells will develop into a tumor and as a result an increase in the number of tumors without an activated H-ras gene occurs. These concepts are based on a number of unproven assumptions. Additional work comparing the molecular and biochemical differences between these different populations of cells will be required to validate or refute this hypothesis. Insight into how these agents enhance the development of tumors in this strain should increase our understanding of chemical carcinogenesis in addition to aiding the interpretation of animal bioassay data. APPENDD( 162 APPENDIX To determine the frequency and mutational profile of H-ras gene activation in spontaneous and chemically induced liver tumors of the B6C3F1 mouse the following experimental protocol was developed. 1. Induce liver tumors in male B6C3F1 mice. 2. Isolate DNA from liver tumor tissue. 3. Amplify, using the polymerase chain reaction technique (PCR), designated regions of the H-ras gene which contain the activating codons 12, 13, 61 and 117. 4. Directly sequence the amplified H-ras gene products. The sections below provide a more detailed description for each of the four procedural steps listed above. 1. Tumor Induction: Male B6C3F1 mice were obtained at the age of 5 weeks from the Charles River Breeding Laboratories (Portage, MI.). Mice were acclimated for 1 week prior to being placed on study. Mice in the benzidine-2H0, phenobarbital and cthroform groups were housed 5 per cage in plastic tubes. Control mice and ciprofibrate treated mice were housed individually in stainless steel cages. Certified rodent chow and municipal tap water were provided ad libitum. Thirty animals were initially placed on study for each of the chemical treatment groups. The dosing regimen for these group is illustrated below. Chemical Dose Route Duration Benzidine-2HC1 120 ppm Drinking water 1 year Phenobarbital 0.05 % Drinking water 1 year Chloroform 200 mg/ kg Corn oil gavage 2X /wk., 1 year Ciprofibrate 0.0125% Diet 2 year 2. DNA Isolation from Liver Tumor Tissue: 163 Genomic DNA was isolated from the frozen tumor tissue as described below. 1. 10. 11. The frozen liver tumor tissue is homogenized in approximately 5 ml of a 10 mM Tris-HCl (pH 7.8), 150 mM NaCl, 2 mM EDTA solution. The appropriate volume of a 10 % SDS solution is added to the tissue homogenate to give a final concentration of 0.5 %. 0.2 volumes of 5.0 M NaClO4 is added, mixed and heated for 10 min. at 60° C. The lysate is cooled to room temperature and extracted with an equal volume of chloroform/isoamylalcohol (25:1). The extract is centrifuge at at 6000 G for 10 min.after which the upper aqueous layer is removed. The DNA is precipitated by the addition of 2.5 volumes of cold ethanol. The DNA is dried, resuspended in 21111 of homogenization buffer and incubated with ribonuclease A (50 ug/ ml) for 1 hr. at 37° C. The above incubation is then followed by an additional incubation with proteinase-k (200 ug/ ml) for 1-2 hr. at 37° C. The incubation mixture is then extracted by successive extractions with an equal volume of phenol; 50:50 mixture of phenol : chloroform/ isoamylalcohol; and chloroform / isoamylalcohol (24:1). Precipitation of the DNA is performed by the addition of 2.5 volumes of cold ethanol. The DNA is dried, resuspended in 10 mM Tris-HCl (pH 7.5), 1 mM EDTA, and quantitated by measuring u.v. absorption at 260 nM. 164 3. Polymerase Chain Reaction Amplification of the H-ras Gene From Liver Tumor DNA: a. Oligonucleotide Primer Preparation. DNA oligonucleotides used as primers for the PCR reaction and DNA sequencing were synthesized using the phosphoramidite method on an Applied Biosystems Model 380A DNA synthesizer. Detritylated primers were purified by HPLC using a 10 u RP-300 reverse phase column. Elution of the primers from the column was achieved with a linear 7-17% acetylnitrile gradient (50 min.,1 ml/ min.) in 0.1 M tetramethylammonium acteate (pH 7.0). A representation of a typical elution profile is provided in Figure 1. HPLC fractions containing the purified primer were pooled, evaporated to dryness, reconstituted with distilled water and purified by ultrafiltration using a Centricon-3. b. Additional Information Concerning PCR Amplification. The PCR reaction was conducted as described in the Materials and Methods section discussed in Chapter 2 of this thesis. Additional information concerning the development of the PCR amplification protocol used in this thesis to amplify genomic tumor DNA is provided below: Initial experiments to amplify the codon 61 region of the H-ras gene from liver tumor DNA resulted in irregular amplification or when amplification did occur low yields were often obtained. These results were confusing because amplification of the liver DNA was concluded under the same conditions which were successful in amplifying a 500 bp region of the lambda genome. After thoughtful analysis two factors were identified as probable sources of the problem. The first problem involved a possible change in the pH and salt concentration of the PCR reaction mixture. Initial development of the PCR oligonucleotide primmer purification involved the pooling and subsequent concentration of HPLC peak fractions containing the purified primers. The eluent used for the HPLC analysis was a 0.1 M tetramethylammonium acetate solution. Concentration of the pooled HPLC fractions resulted in an approximate 50 to 100 fold increase in the salt concentration of the primer solution. Addition of the primer solution to the PCR reaction volume would increase the salt concentration by 5 to 10 fold. Also, the addition of the concentrated acetate solution may have changed the pH of the PCR reaction. Both the increase in the salt concentration and change in pH may have 165 Figure 1: Representative chromatogram of HPLC purification for DNA oligonucleotides primers used in PCR amplification and DNA sequencing. Parameters used for the analysis are discribed in the Materials and Methods section within Chapter 2. 166 Figure 1: 18 met Absorption 254 nm Distance [cm] 167 adversely inhibited the Taq polymerase. To eliminate this problem the high salt concentration of the primmer solutions was removed, prior to use in the PCR reaction, by ultrafiltration. The second problem involved the accessibility of the Taq enzyme to the H-ras gene sequence. Purification of genomic DNA at high concentrations results in a very viscous suspension. The high viscosity of this suspension leads to poor solubility when a small aliquot is added to to the PCR reaction volume. Because of this, it was rationalized that the the H-ras gene may be poorly accessible to the Taq enzyme. To alleviate this potential problem an experiment was conducted to determine whether preincubation of the genomic DNA with the endonuclease Bam H1 would increase the efficiency of the PCR reaction. It was rationalized that the digestion of the genomic DNA with the endonuclease would break down the long genomic fragments facilitating the dispersion of the DNA in the reaction solution and thereby increase the template accessibility to the Taq enzyme. The effectiveness of the Barn H1 preincubation on PCR amplification effeciency is illustrated in Figure 2. 168 Figure 2: Agarose electrophoretic gel analysis demonstrating the utility of Bam H1 digestion on increasing PCR amplification efficiency. PCR amplifications were conducted as described in the Materials and Methods section of Chapter 2. Lane 1: Hind III digest of lambda DNA Lane 2: 1.75 pg of normal liver DNA (undigested) was subjected to PCR amplification of an 80 bp region within exon 2 of the H-ras gene. Lane 3: 5.25 ug of normal liver DNA (undigested) was subjected to PCR amplification of an 80 bp region within exon 2 of the H-ras gene. Lane 4: 1.75 pg of normal liver DNA was digested with Barn H1 endonuclease prior to PCR amplification of an 80 bp region within exon 2 of the H-ras gene. 169 Figure 3: Agarose electrophoretic gel analysis from a representative PCR amplification of a 110 bp region within the H-ras gene from liver tumor tissue. Total genomic DNA, isolated from independent liver tumors induced with ciprofibrate, was subjected to PCR amplification as described in the Materials and Methods within Chapter 2. The sequence of the amplified 110 bp region is illustrated in Figure 1 of Chapter 2. Lane 1 is a Hind III digest of lambda DNA; Lanes 2-14 are the ciprofibrate samples. 170 (a) 117-A Figure 4: Representaive DNA sequence analysis of the H-ras gene codon 12, 13 and 117 regions. PCR amplified regions of exon 1 and 3 within the H-ras gene (illustrated in Figure 1 of Chapter 2) were sequenced as described in the Materials and Methods section of Chapter 2. (a) representative sequence of the H-ras gene encompassing codons 12 and 13. The DNA derived for this analysis was from a spontaneous liver tumor. (b) representative sequence of the H-ras gene encompassing codon 117. The DNA derived for this analysis was also derived from a spontaneous liver tumor. "Illilllllllllllllllll