\ i-‘i‘té: .2354," )‘v\"$: ‘~. .3?wa ' V. ”lg ’ '13, ’v “y K "I ';'. I I.) ‘ N};Il“ -' , ' . I g 1.3:“ J’- 3“!) ‘ l "'14 §I:~..N .h- "k ‘ .16 11%..» n A 0::3; ‘I\ ' ~13; " "fig-2‘ . .. .9331: 2:39.?» A 32.1 .,. 4)- 3.: w ' gawk/421:3...“ .3; » )- :t; \‘3hakkvp‘wifi‘41 3:» ‘1 . \ :‘v: s' . .{fi' \‘I‘V- ; “ch!“ )7? ix:‘..' x)“: “is {31‘ 3:3: {53?}. ‘3' $5? . 23 by: " $3.1m . é; (“max-m 5.7.. ..§\,'T , 2”?”2 ‘f/‘J' :4.’ 2r .3 I, . 'l. '3’ l I .1, adv" 911-; ~.. 1"]‘0” 22:81:," . 1 li‘ ‘5 14$ I ”.3; 1‘1.) .‘.’ it" '1 {id #4 V0 Kin w’. :Jdfik IV (.556. ‘ I ‘ 9“ . " — ii: 31 «f: » ,_- i“ ;\ x ’ _ '3. a. .Q‘-“ ..,. ' '3. 4““ _\3 4‘ M‘ 3:1! #1)"! 4:} -f vjééf. in." :Q‘. lf .3. “: L-C ‘ ',.u 1’3.“ 5‘1, - a Y1 fi . P ‘5 ,. ,I . . xfij ¥ , '.“' o' u. 1- -\:\'V\1‘ .' '. z;l.}::.'\ ,.\1 ”(3.: 1.3:.” u‘, ' ‘1 ”a ‘ a .- K. g. .. .57. . 3.," 2;: l’" ' J- ”J. qw' 1:: ‘1‘ ”I 11 .7 . V. ‘33“. . ‘11:. 1,. Vbivxfiagfifc a .}(" Eitf‘cfig‘. 'v v? '4 ‘a m .g. ‘3 IV ”‘9‘“. > 'nn".:.' ‘7 2:? "} I} “If"fi” fay} PW”! I": I ':"'_ , )' fig-.1». 5f”. ‘Qr'flj'l' I’ll}; ‘A:-I ”3,4 " ;‘i :93...“ 3 ' ”v.1. ‘ 3"/JA'.‘£/.21 -"-V-.|rr,_,,. tidal": -;ym...-. ' .‘v l HIGAN STATE UNIVERSITY LIBRARI WITHIN!!! l nu lHll mmmfifiu 3 12'93 00551 0767 LIBRARY ‘ Michigan State University '94ng 2"i(9% This is to certify that the dissertation entitled Malignant Transformation Induced in Human Fibroblasts by Expression of a Transfected T24 Hara§_0ncogene: Requirement for an Infinite Lifespan presented by Peter J. Hurlin has been accepted towards fulfillment of the requirements for Ph.D. Biochemistry degree in ' (7M BEAM/11K 4 Major professor Date 7/ 1/88 MS” it an Aflinnan'vc Action/Eq ual Opportunity Institution 0-12771 MSU LIBRARIES .-_ RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below. MALIGNANT TRANSFORMATION INDUCED IN HUMAN FIBROBLASTS BY EXPRESSION OF A TRANSFECTED T24 H-ng ONCOGENE: REQUIREMENT FOR AN INFINITE LIFESPAN by Peter J. Hurlin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1988 l 1-... ‘5" r-c-O _,.. 4,“: ~o~ ( L... x',’ a). ' II {(- .‘ It ABSTRACT MALIGNANT TRANSFORMATION INDUCED IN HUMAN FIBROBLASTS BY EXPRESSION OF A TRANSFECTED T24 H-[gg ONCOGENE: REQUIREMENT OF AN INFINITE LIFESPAN by Peter J. Hurlin This study was undertaken to identify the phenotypes induced in human fibroblasts by expression of the T24 H-ras oncogene, one of several members of the gas family of genes freqently detected in human tumors. When the T24 H-ras oncogene, in a plasmid containing two sets of enhancer sequences, was transfected into early passage, finite lifespan human fibroblasts, colonies of morphologically transformed cells were generated. Progeny of these cells expressed T24 H-ras p21, formed foci, and developed colonies in soft agar. No such transformation was observed following transfection with the normal H-ras gene. The majority of transformed populations reverted to the normal phenotype, which correlated with loss of expression of T24 H-ra; p21, even though the DNA persisted. The stably transformed cells exhibited a finite lifespan and were not tumorigenic. To determine whether the finite lifespan of these cells was a factor in preventing tumor formation, the T24 H-ras oncogene was transfected into an infinite lifespan human fibroblast strain, MSU-l.l, generated in this laboratory following transfection with a v—myc oncogene. Foci of morphologically transformed cells developed, and progeny of these cells expressed the v-myc p110 and T24 H-L§§ p21, formed colonies in soft agar, proliferated rapidly in serum-free medium containing 0.1 mM calcium, which does not support growth of the parental cell line, and formed progressively growing malignant fibrosarcomas in athymic mice. No such transformation was observed with the control plasmid. Transfection of additional infinite lifespan human fibroblast cell lines with the T24 H- m oncogene also resulted in malignant transformation. The results demonstrate the potent ability of the T24 H-gas oncogene to induce in human fibroblasts characteristics of tumor cells and the fundamental and complementary role that the infinite lifespan phenotype plays in T24 H- ;gs oncogene-induced malignant transformation. ACKNOWLEDGEMENT I would like to thank.my graduate advisor and research director, Dr. J. Justin McCormick, for sharing with me his broad knowledge of cancer, and his many ideas and perspectives that have shaped this thesis. I also owe a special thanks to Dr. Veronica Maher for her direction and for her critical review of the written manuscripts describing my work. 1 also would like to express my graditude to my parents. Shirley and Kenneth Hurlin, whose friendship, guidance and love have been constant throughout this endeavor. Finally, the love, support and smile of Patricia Pillsbury, my'companion through the last two years, is gratefully acknowledged. iii TABLE OF CONTENTS Page LIST OF TABLES ................................................... vii LIST OF FIGURES .................................................. viii INTRODUCTION ..................................................... 1 CHAPTER I. LITERATURE REVIEW A. Role of oncogenes in tumorigenesis ..................... 8 l. The genetic basis of cancer ......................... 8 2. Identification of oncogenes in the genomes of acutely transforming retroviruses ................... 9 3. Identification of cellular oncogenes in human and animal tumors ....................................... 11 a. Oncogenes identified in tumors by the NIH 3T3 transformation assay ............................. 11 b. Identification of oncogenes by genetic rearrangement .................................... 13 i. Activation by proviral insertion ............ 13 ii. Activation by chromosomal translocation ..... 14 iii. Activation by amplification ................. 15 4. Function of oncogene encoded products in tumorigenesis ...................................... 16 a. Growth factors and growth factor receptors ...... 16 b. Intracellular modulators of cell growth ......... 18 c. Nuclear proteins involved in contolling gene expression ...................................... 18 B. Biochemical properties of La; ......................... 20 l. The gas genes ...................................... 20 a. Structural features of gas genes ................ 20 b. Mechanism of activation ......................... 22 2. The [as proteins ................................... 23 a. Structure of gas proteins .............................. 23 b. GTP binding and hydrolyzing activities .......... 25 c. Interaction with a putative effector protein.... 26 d. Involvement in the control of cell proliferation 27 3. Biological properties associated with expression of gas oncogenes in mammalian cells ................... 28 a. Biological properties induced in rodent fibroblasts ..................................... 28 b. Biological consequences of expression of [as oncogenes in transgenic mice .................... 30 c. Use of human cells to study the biological effects of gas oncgene expression ............... 32 References ............................................ 34 iv CHAPTER II. MORPHOLOGICAL TRANSFORMATION, FOCUS FORMATION AND ANCHORAGE INDEPENDENCE INDUCED IN DIPLOID HUMAN FIBROBLASTS BY EXPRESSION OF A TRANSFECTED H-ggs ONCOGENE Summary .................................................... 48 Introduction ............................................... 49 Materials and Methods Cells and culture conditions ............................. 51 Plasmids ................................................. 51 DNA transfection ......................................... 51 Selection of G418 resistance ............................. 52 Assay for focus formation ................................ 52 Assay for anchorage independence ......................... 52 Southern blot analysis ................................... 53 RNA analysis ............................................. 53 Detection of p21 ......................................... 53 Results Frequency of transfection to G418 resistance ............. 55 Transformation to altered morphology ..................... 55 Transformation to focus formation ........................ 58 Induction of anchorage independence ...................... 58 Analysis for the presence of transfected sequences ....... 61 Expression of the T24 H-ras oncogene in transformed cells 64 Tumorigenicity ........................................... 69 Discussion .................................................. 70 Acknowledgements ............................................ 73 References .................................................. 74 CHAPTER III. MALIGNANT TRANSFORMATION OF A HUMAN FIBROBLAST CELL STRAIN, MSU-1.l, BY EXPRESSION OF A TRANSFECTED T24 H-Lgs ONCOGENE Summary ..................................................... 80 Introduction ................................................ 81 Methods and Materials Cells and culture conditions ............................. 84 Plasmids and DNA transfection ............................ 84 Assay for growth factor independence ..................... 85 Assay for anchorage independence ......................... 85 Tumorigenicity assays .................................... 85 DNA analysis ............................................. 86 Analysis of gas p21 and v-myg p110 ....................... 86 Karyotype analysis ....................................... 87 Results Transformation to focus-formation by the T24 H-ras oncogene ................................................. 88 Growth of MSU-l.l-T24 fibroblasts in serum-free medium... 88 Anchorage independence of MSU-1.l-T24 cell strains ....... 90 Tumorigenicity of MSU-1.l-T24 cell strains ............... 90 Presence of T24 H-ras and v-myg oncogenes ................ 94 Expression of T24 H-ras and v-myg oncogenes .............. 95 Karyotype analysis of MSU-1.l-T24 cell strains ........... 99 Tumorigenic transformation of infinite lifespan human fibroblasts derived from other cell lines ................ 100 Discussion .................................................. 102 Acknowledgments ............................................. 106 References .................................................. 107 APPENDIX A ....................................................... 111 APPENDIX B ....................................................... 112 APPENDIX C ....................................................... 113 vi LIST OF TABLES Chapter II Page 1. Frequency of transformation to altered cellular morphology and focus formation .............................................. 56 Chapter III 1. Growth properties and tumorigenicity of MSU-1.1-T24 cell strains .......................................................... 93 vii LIST OF FIGURES Chapter I Page 1. Comparative amino acid sequence of [as proteins ............... 24 Chapter II 1. Comparison of normal and transformed morphologies ............ 57 2. The focus-forming ability of pHOGTl-transfected fibroblasts ...................................................... 59 3. Southern blot analysis of cellular DNA from morphologically transformed cells obtained following transfection with pH06T1, and there morphologically normal progeny ............................. 62 4. Analysis of [as expression ................................... 65 5. H-Las RNA expression ......................................... 67 Chapter III 1. Morphology of LGI fibroblasts, MSU-1.l fibroblasts and MSU-1.1-T24 fibroblasts .......................................... 89 2. Histology showing a poorly-differentiated human fibrosarcoma formed in an athymic mouse by MSU-1.1-T24 strain 3 cells ......... 91 3. Southern blot analysis of plasmid pH06Tl-derived T24 H-Las sequences in MSU-1.l-T24 cell strains ...................... 95 4. Expression of T24 H-Las p21 in MSU—l.l-T24 cell strains ...... 96 5. Expression of v-myg encoded p110 in MSU-1.1-T24 cells strains .......................................................... 98 APPENDIX A Map of plasmid pH06T1 ............................................ 111 APPENDIX B Athymic mouse with tumor formed by subcutaneous injection of MSU-l.l-T24 strain 3 ............................................. 112 APPENDIX C 1. Comparison of gas p21 expression in focus-derived, anchorage independent colony-derived and tumor-derived MSU-1.1-T24 cell strains .......................................................... 114 2. Southern blot analysis of MC29 viral sequences in MSU-1.1-T24 cell strains ..................................................... 116 3. Analysis of 3;; p21 expression in KMST6-T24, GM0637-T24 and SUSMl-T24 cell strains ........................................... 118 viii Introduction Malignant transformation of cultured mammalian cells is generally recognized as a multistep process (Barrett and Ts’o, 1978; Thommasson and DeMars, 1982; Perez-Rodrigez et al. 1982: Newbold et al., 1982; Land et al., 1983; Ruley, 1983), as is in vivo carcinogenesis (Peto, 1977; Doll, 1978). Overwhelming evidence exists indicating that agents which damage DNA are capable of mediating steps in the pathway leading to malignancies in vivo and transformation in vitro. The first clues to the identification of potential genetic targets for such agents came from studies on RNA tumor viruses. Such studies showed that cellular genes, transduced by RNA tumor viruses, were responsible for greatly enhancing the ability of these viruses to cause tumors in host animals (Varmus, 1982). Studies on RNA tumor viruses converged with DNA transfection 'studies early in this decade resulting in the discovery that human tumors contained mutated versions of cellular genes, which were homologous to some of the cellular transforming genes of RNA tumor viruses (Bishop, 1983; Varmus, 1984). Cellular oncogenes were first detected in the DNA of tumors by their ability to induce characteristics of tumor cells, for example focus formation, when transfected into mouse NIH3T3 cells (Shih et al., 1979). The oncogenes most frequently detected in human tumors using the NIH3T3 transfection assay are homologous to the transforming gas genes of Harvey and Kirsten murine sarcoma viruses, and members of a family of cellular genes which have been designated H-gas, K-gas and N- gas (Varmus, 1984). Sequence analysis of these oncogenes indicated that 2 point mutations occurring at or near codons 12 (Tabin et al., 1982; Reddy et al., 1982; Taparowski et al., 1982) or 61 (Taparowski et al., 1983; Yuasa et al., 1983) were responsible for their ability to transform NIH 3T3 cells. The detection of gas oncogenes in approximately 10-15%.of human tumors (Barbacid, 1987), together with the known association between carcinogen exposure and the generation of point mutations (Ames, 1979), provides strong circumstantial evidence of a role for gas oncogenes in the development of human malignancies. Our understanding of the biological properties associated with expression of gas_oncogenes has come primarily from transfection studies using rodent fibroblasts. Early transfection experiments using finite lifespan diploid rodent fibroblasts indicated that gas oncogenes could induce several phenotypes characteristic of tumor cells, including morphological transformation, focus formation and anchorage independence, but not an infinite lifespan in culture or tumorigenicity (Land et al., 1983; Ruley, 1983; Newbold and Overell, 1983). However, co-transfection of such cells with an H-gas oncogene and any one of a number of genes thought to cause induction of an infinite replicative lifespan in culture resulted not only in morphological transformation, focus formation and anchorage independence, but also in an infinite lifespan in culture and tumorigenicity (Land et al., 1983; Ruley, 1983). These results, along with results showing that transfection of ms oncogenes into rodent fibroblasts already exhibiting an infinite lifespan caused tumorigenic transformation (Land at al., 1983; Overell and Newbold, 1983), led to the hypothesis that oncogenes that induce the infinite lifespan phenotype act in a complementary or cooperative fashion with gas oncogenes in a two step model of in vitro carcinogenesis. 3 Although the use of rodent fibroblast transformation systems has been instrumental in the identification and preliminary characterization of oncogenes, it is not clear whether the results obtained using rodent fibroblasts are directly applicable to cancer development in humans. This is because exposure to various carcinogens (Barrett and Ts’o, 1978), transfection of cloned oncogenes (Land at al., 1983; Ruley, 1983; Newbold and Overell, 1983), or direct microinjection of oncogene-encoded products (Feramisco et al., 1984; Stacy et al., 1984) can cause transformation of rodent fibroblasts in culture, but attempts to reproduce such results using human fibroblasts for the most part have failed (DiPaolo, 1983; Sager et al., 1983; Spandidos et al., 1984; Feramisco et al., 1984; Stacy et al., 1984). The failure of human fibroblasts to transform as readily as rodent fibroblasts in culture may result from a fundamental biological difference between the species, or may be related to the fact that, unlike human fibroblast cell lines, many rodent fibroblast cell lines can spontaneously transform into cells that exhibit characteristics of tumor cells (Heidelberger, 1973). Because of this latter property of rodent fibroblasts, one cannot dismiss the possibility that some transformed characteristics attributed to the mutagenic effects of a carcinogen, or expression of a transfected oncogene may actually be the result of, or related to some spontaneous transformation event(s). The goal of this thesis was to define the biological properties induced in human fibroblasts following transfection of these cells with the T24 H-gas oncogene, an H-gas oncogene cloned from a human bladder carcinoma. Results from earlier studies by others who tried to accomplish this same goal suggested that human fibroblasts in culture are resistant to 4 transformation by gas oncogenes (Sager et al., 1983; Spandidos, 1985). However, the considerable circumstantial evidence involving this family of oncogenes in the development of human malignancies, as discussed above, encouraged me to carry out a more comprehensive examination of the problem. Transfection studies with the T24 H-gas oncogene cloned into various vectors were carried out using both finite lifespan, and infinite lifespan human fibroblasts in order to permit comparison of the results to those of gas transfection studies that used comparable rodent fibroblast cell lines as recipients. Chapter I of the thesis reviews the literature that provides the conceptual framework of our understanding of how oncogenes can be activated and how they may be involved in the induction of cancer and reviews the literature directly concerning gas oncogenes. Chapter II is primarily a manuscript by P.J. Hurlin, D.G. Fry, V.M. Maher and J.J. McCormick that was published in the November, 1987 issue of Cancer Research 51, 5752-5757, which describes my work showing that transfection of the T24 H-ras oncogene into diploid human fibroblasts induces several characteristics of tumor cells. The published manuscript has been modified in Chapter II to include data showing that cells stably transformed following transfection of the T24-H-gas oncogene are not tumorigenic. These latter data were published in a review article by D.G. Fry, P.J. Hurlin, V.M. Maher and J.J. McCormick which appeared in a special issue of Mutation Research (vol. 199, 341-351, 1988) on the transformation of human fibroblasts. Chapter III is a manuscript by P.J Hurlin, V.M. Maher and J.J. McCormick submitted in May, 1988 for consideration for publication in the Proceedings of the National Academy 5 of Sciences USA, which describes my work showing that the T24 H-ras oncogene induces malignant transformation when it is transfected into infinite lifespan human fibroblasts. Data relevant to this study that were not shown in the submitted manuscript are shown in Appendices B and C. Appendix A shows maps of plasmids important to these studies. REFERENCES Ames, B.N. (1979) Identiying environmental chemicals causing mutations and cancer. Science 885, 587-593. Barbacid, M. (1987) gas genes. Ann. Rev. Biochem. 88, 779-827. Barrett, J.C. and Ts’o, P.0.P. (1978) Evidence for the progressive nature of neoplastic transformation in vitro. Proc. Natl. Acad. Sci. USA 18, 3761-3765. Bishop, J.M. (1983) Cellular oncogenes and retroviruses. Ann. Rev. Biochem., 52, 301-354. DiPaolo, J.A. (1984) Relative difficulties in transforming human and animal cells lg vitro. J. Natl. Cancer Inst. 18, 3-8. Doll, R. (1978) An epidemiological perspective of the biology of cancer. Cancer Res. 88, 3573-3583. Feramisco, J.R., Gross, M., Kamata, T., Rosenberg, M. and Sweet, R.N. (1984) Microinjection of the oncogene form of the human H-gas (T-24) protein results in rapid proliferation of quiescent cells. Cell 88, 109- 117. Heidelberger, C. (1973) Chemical carcinogenesis in culture. Advances in Cancer Research, Vol. 18, 318-365, Academic Press, Inc., New York and London. Land, H., Parada, L.F. and Weinberg, R.A. (1983) Tumorigenic conversion of primary embryo fibroblasts requires at least two cooperating oncogenes. Nature 885, 596-602. Land, H., Parada, L.F. and Weinberg, R.A. (1983) Cellular oncogenes and multistep carcinogenesis. Science 888, 771-778. Newbold, R.F. and Overell, R.N. Fibroblast immortality is a prerequisite for transformation by EJ Ha-gas oncogene. (1983) Nature 895, 648-651. Perez-Rodriguez, R., Franchi, A., Deys, B.F. and Pouyssegur, J. (1982) Evidence that hamster fibroblast tumors emerge in nude mice through a process of two in vivo selections leading to growth factor "relaxation" and to immune resistance. Int. J. Cancer 88, 309-314. Peto, R. (1977) The origins of human cancer. Book C, Hiatt, H.H., Natson, J.D. and Hinsten, J.A. (Eds.), Cold Spring Harbor Laboratory, New York, pp. 1403-1437. Reddy, E.P., Reynolds, R.K., Santos, E. and Barbacid, M. (1982) A point mutation is responsible for the acquisistion of transforming properties by the T24 human bladder carcinoma oncogene. Nature, 888, 149-152. 7 Ruley, H.E. (1983) Adenovirus early region 1A enables viral and cellular transforming agents to transform primary cells in culture. Nature 888, 602-606. Sager, R., Tanaka, K., Lau, C.C., Ebina, Y. and Anisowicz, A. (1983) Resistance of human cells to tumorigenesis induced by cloned transforming genes. Proc. Natl. Acad. Sci. USA 88, 7601-7605. Shih, C., Shilo, B.Z., Goldfarb, M.P., Dannenberg, A, and Weinber, R.A. (1979) Passage of phenotypes of chemically transformed cells via transfection of DNA and chromatin. Proc. Natl. Acad. Sci. USA 18, 5714- 5718. Spandidos, D. (1985) Mechanism of carcinogenesis: The role of oncogenes, transcriptional enhancers and growth factors. Anticancer Res. 8, 485-498. Stacy, D.W. and Kung, H-. F. (1984) Transformation of NIH 3T3 cells by microinjection of Ha-gas p21 protein. Nature 818, 508-510. Tabin, C.J., Bradley, S.N., Bargman, C.I., Weinberg, R.A., Papageorge, A.G., Scolnick, E.M., Dhar, R., Lowy, D.R. and Chang, E.H. (1982) Mechanism of activation of a human oncogene. Nature 888, 143-149. Taparowski, E., Shimizu, K., Goldfarb, M. and Wigler, M. (1983) Cell 88, 581-586. Taparowski, E., Suard, Y., Fasano, 0., Shimizu, K., Goldfarb, M. and Wigler, M. (1982) Activation of the T24 bladder carcinoma transforming gene is linked to a single amino acid change. (1982) Nature 888, 762- 765. Thommasen, D. G. and DeMars, R. (1982) Clonal analysis of the stepwise appearance of anchorage independence and tumorigenicity in CAK, a permanent line of mouse cells. Cancer res. 88, 4054-4063. Varmus, H.E. (1982) Form and function of retroviral proviruses. Science 818, 812-820. Varmus, H.E. (1984) The molecular genetics of cellular oncogenes. Ann. Rev. Genet. 18, 553-512. Yuasa, Y., Srivastava, S.K., Dunn, C.Y., Rhim, J.S., Reddy, E.P. and Aaronson, S.A. (1983) Acquisition of transforming properties by alternative point mutations within c-bas/has human proto-oncogene» Nature 888, 775-779. CHAPTER I LITERATURE REVIEH A. Role of Oncogenes in Tumorigenesis 1. The Genetic Basis of Cancer Cancer incidence data indicate that human populations living in different geographical regions of the world tend to develop different kinds of cancer (Waterhouse, 1976). For example, the inhabitants of third world countries often have a high incidence of liver cancer, those of western nations have a high incidence of cancer of the breast and colon, and those living in Japan have a high incidence of stomach cancer. Migrant populations tend to take on the pattern of cancer that is characteristic of their new homes (Waterhouse, 1976), suggesting that such variation in cancer incidence is due to diet, customs, or what can be generally considered environmental factors. It is now well established that the two most common cancers in the western world, skin cancer and lung cancer, are caused primarily by known environmental factors, i.e., sunlight exposure and cigarette smoking respectivelyu Components of cigarette smoke and ultraviolet irradiations, like most carcinogenes interact with DNA and are mutagenic (Haseltine, 1983). Because skin, lung and most other types of cancer are strongly age dependent, it has been hypothesized that cancer induction must be a multistep process, and that the accumulation of critical mutations, resulting from carcinogen exposure or otherwise, over the life of an individual is part of this process (Doll, 1978; Cairns, 1981; Moogavkar and Knudson, 1981). Support for this hypothesis has come from a wide 9 range of studies including those on the induction of cancer in animals by chemical carcinogens (Berenblum, 1940; Berenblum and Shubik, 1947), studies on the relationship between carcinogen exposure and the induction of mutation and cell transformation of cells in culture (Barrertt and Ts’o, 1978; Spandidos and Siminovitch, 1978), and studies on the increased mutagenicity of carcinogens in cells derived from individuals who have a hereditary predisposition to cancer (Maher et al. 1976). There is new experimental evidence to show that point mutations and deletions, gene amplification and chromosomal translocation- -events that are known or suspected to be induced by carcinogens- -can cause the "activation" of specific cellular genes, referred to as "proto-oncogenes", into oncogenes whose protein products contribute to the malignant phenotype. 2. Identification of Oncogenes in the Genomes of Acutely Transforming Retroviruses It was hypothesized by Temin in 1964 that as part of the lifecycle of RNA tumor viruses, RNA acted as a template for the synthesis of a DNA provirus which integrated into the host cells genome. The discovery of the reverse transcriptase enzyme by Temin and Baltimore in 1970 explained how the DNA-provirus mechanism could work. Experiments by Neiman (1972), showing the presence of proviral DNA in the genome of infected cells, provided more formal proof of the hypothesis. These discoveries led to explanations for the profound transforming properties of some RNA tumor viruses such as Rous sarcoma virus (Rous, 1911). It was proposed by Temin (1976) that the genome of acutely transforming retroviruses evolved, in part, from normal cellular sequences that had been previously mutated in the host cell by exposure to carcinogens. 10 A similar hypothesis, the "viral oncogene hypothesis", proposed by Huebner and Todaro in 1969, stated that the cells of all or most vertebrate species contain genomes of potentially oncogenic RNA tumor viruses that are transmitted vertically from parent to offspring. In a modified version of this hypothesis (Todaro and Huebner, 1972) it was proposed that the occurrence of cancer may be determined by the activation of normally repressed viral genes by, perhaps, exposure to carcinogenic agents or the normal processes of aging. Experiments by Neiman (1972) and later Spector (1978) showing that sequences within the acutely transforming Rous chicken sarcoma virus hybridized ‘with the DNA, of uninfected cells appeared to support this hypothesis. However, it is now known that the life cycle of RNA tumor viruses predisposes them to rare genetic events that lead to the incorporation in their genome a group of cellular genes that provide a selective advantage for proliferation to cells that are infected by such transforming viruses. Ultimately, this is manifested as malignant neoplasia. Although it is not true, as originally hypothesised by Huebner and Todaro (1969), that retroviral genomes are present in all of the cells of any given host species, it is true that specific genes (proto-oncogenes) which have the potential to contribute to the malignant phenotype do exist in all cells. This concept, and the concept, put forth by both Huebner and Todaro (1969) and Temin (1976), that carcinogen induced mutations can be responsible for activating the transforming potential in such cellular genes are now well established. Since the first v-oncogene, v-sgs, was identified in 1976 (Stehelin et al., 1976) there have been more than 20 other v-oncogenes identified in the genomes of RNA tumor viruses isolated from tumors, mostly sarcomas, 11 formed in a number of different vertebrate species. Each of these oncogenes has been found to be homologous to a cellular counterpart in the genome of the viruses host species. Such cellular counterparts or proto- oncogenes exist in evolutionarily distant eukaryotes including yeast and drosophila and are sometimes regulated in a cell cycle or lineage-specific manner, suggesting that they might play a central role in processes that control the normal cellular growth and proliferation, processes that are altered in cancer cells. 3. Identification of Cellular Oncogenes in Human and Animal Tumors There are many possible mechanisms by which cellular proto-oncogenes can be activated into oncogenes. These include overexpression, inappropriate or ectopic expression (e.g., expression in cell type that it is normally not expressed in), unscheduled expression, altered biochemical properties caused by structural alterations within coding sequences, or some combination of these mechanisms. In most cases, putative cellular oncogenes have been detected in human and animal tumors by the identification of’ genes that possess one or more of these properties, as well as by exibiting homology to recognized retroviral oncogenes. a. Oncogenes Identified in Tumors by the NIH3T3 Transformation Assay Experiments by Anderson et al. (1979) demonstrated that DNA fragments of acutely transforming retroviruses, that carried putative v- oncogenes, were capable of causing cellular transformation when transfected into normal cells. These experiments suggested to Weinberg and his colleagues that if cellular counterparts (c-oncogenes) existed in 12 chemically transformed cells, they too should be capable of exerting such dominant transforming effects on normal cells. Experiments by Shih et al. (1979), demonstrated that when DNA, isolated from chemically transformed mouse cells, was transfected into non-transformed mouse NIH 3T3 cells using the the calcium phosphate transfection technique developed by Graham and Van der Eb (1973), foci of morphologically transformed cells were generated. When DNA, isolated from normal mouse cells, was transfected into the same recipient cells, it did not cause such transformation. Although there was no formal proof, it was concluded that the induced transformation was the result of the transfer of cellular alleles (oncogenes) that had their transforming potential activated by carcinogen induced mutations (Shih., 1979). The existence of such transforming sequences has since been demonstrated in the DNA of a large number of different human tumor cell lines and human tumor biopsy material. The identification of tumor-derived sequences that caused fecus formation and malignant transformation when transfected into mouse NIH 3T3 cells was first achieved by four separate groups (Parada et al., 1982; Santos et al., 1982; Der et al., 1982; Goldfarb et al., 1982), using DNA derived from human bladder carcinoma cell lines. By using molecular probes developed from v-oncogenes, each of these groups discovered that the sequences responsible for transformation in each case were homologous to the v-oncogene, v-H-gas, of the Harvey strain of murine sarcoma virus. Literature concerning gas genes and their products is reviewed in detail in later sections. Several other putative oncogenes have been detected in human tumor cell DNA by the NIH3T3 transformation assay (Padhy et al., 1982; Cooper 13 et al., 1984; Padua et al., 1984; Shimizu et al., 1985; Eva and Aaronson, 1985; Martin—Zanca et al., 1986; Fusco et al., 1987). For at least two of these oncogenes the putative mechanism of activation has been identified. The gag oncogene, isolated from a rat neuroblastoma, codes for a transmembrane, receptor-like protein and shows a 50% sequence identity with the human epidermal growth factor receptor gene (Schechter et al., 1984) appears to be activated by point mutations (Bargman et_al., 1986). The gag oncogene, isolated from a human colon carcinoma, is a fusion gene consisting of sequences derived from the tropomyosin gene and a gene that codes for a protein tyrosine kinase (Martin-Zanca et al., 1986). This fusion gene appears to have been generated by a mechanism involving chromosomal translocation (Martin-Zanca et al., 1986). b. Identification of Oncogenes by Genetic Rearrangements i. Activation by Proviral Insertion Retroviruses that do not have oncogenes may nevertheless be involved in tumor induction. This occurs by means of insertional mutagenesis as was first shown in the study of chicken lymphomas (Hayward et al., 1981). In these tumors, the cellular myc gene has been activated by insertions of retroviral DNA upstream, or within the gene (Hayward,et al., 1981; Payne et al., 1981; Neel et al., 1981). Altered transcriptional regulation of c-myc is generally thought to be one of several steps in tumorigenesis (Thompson and Neiman, 1987). Such altered regulation and/or overexpression of a previously regulated or silent gene results because the integrated viral DNA provides a strong promoter from which transcription of the cellular gene is initiated and/or a transcriptional enhancer that stimulates increased transcription from the cellular gene’s 14 own promoter (Bishop, 1987). Because proviral integration is thought to occur randomly throughout the host cells genome (Varmus, 1983), commonly interrupted cellular domains identified in tumor cells are likely to represent sites that are functionally significant to tumor induction. In addition to c-mys, c-mas (Gattoni-Celli et al. 1983; Cohen et al., 1983;) and c-ag1)_8 (Fung et al., 1983), which has been identified as the epidermal growth factor receptor gene (Downward et al. 1984), are frequently activated in animal tumors by proviral insertional mutagenesis. Further support for a role in tumor induction by these genes is based on the fact that each is homologous to v-oncogenes (Varmus, 1984). Other genetic loci that do not show homology to v-oncogenes have also been identified as putative oncogenes based on the observance of repeated adjacent retroviral insertion (Varmus, 1984). ii. Activation by Chromosomal Translocations Although most human tumors fail to show specific chromosomal translocations, there are highly regular translocations associated with some tumor types, most notably chronic myelogenous leukemia (CML), acute lymphoblastic leukemia (ALL), and several B-cell malignancies (Sandberg et al. 1980). In approximately 95% of CML patients and 20% of ALL patients (Sandberg, 1980), there exists a cytogenetic marker, the Philadelphia chromosome, that is generated by a reciprocal translocation between chromosomes 22 and 9 (Nowell and Hungerford, 1960). The presence of the Philadelphia chromosome in such patients has important prognostic and diagnostic value (Sandberg, 1980). The c-a81 gene, a homologue of the v-a81 oncogene that codes for a protein tyrosine kinase (Davis et al. 1985), has been localized to the portion of chromosome 9 that contributes 15 to the Philadelphia chromosome (de Klein et al., 1982). The c-a81 gene is translocated into a region of chromosme 22, designated the breakpoint cluster region (gag), and the biochemical consequence of this molecular rearrangement is the production of an abnormal fusion protein, gag-aal p210, with enhanced protein kinase activity (Konopka and Witte, 1985; Chan et al. 1987; Kurzrock et al. 1987). Whereas the translocations that characterize the formation of the Philadelphia chromosome in CML and ALL apparently activate the a81 gene by altering the biochemical properties of its product, translocations characteristic of a number of B-cell malignancies appear to activate the c-mya gene by altering its normal transcriptioal regulation (Leder et al., 1983, Marcu et al., 1983). The specific translocations existing in a large number of B-cell tumor lines join the c-mys gene on chromosome 8 with the an immunoglobulin locus on one of three different chromosomes (2, 14 and 22) (Croce et al., 1979; Erickson et al., 1982; Malcolm et al., 1982). Altered transcriptional regulation of the c-mys gene (overexpression and/or unscheduled expression), caused by the influence of adjacent transcriptional regulatory elements at the immunoglobulin loci, appears to be the mechanism of activation of the c-mgs gene in such B-cell malignancies (Leder, 1983). iii. Activation by Amplification Several cellular genes, that are homologous to v-oncogenes, have been identified in some tumors and tumor-derived cell lines in multiple copies. For example, the mys proto-oncogene is amplified 30 to 50-fold in the human promyelocytic leukemia cell line HL-60 (Dalla-Favera 1982), 16 and in small cell carcinoma-derived cell lines (Little et al., 1983). The k-gas proto-oncogene is amplified 3- to 5-fold in a human colon carcinoma cell line (McCoy et al., 1983), and a member of the mys family of genes designated N-mys is amplified in many human neuroblastomas (Schwab et al., 1983). Such amplification has been found to correlate with advanced disease stage (Brodeur et al., 1984). In these cases, the increased gene copy number correlates with an increased level of transcription. 4. Function of Oncogene Encoded Products in Tumorigenesis For the most part, the physiological function of the products of oncogenes is unknown. However, the preliminary characterization of oncogene encoded products has revealed that they all appear to share properties with various gene products that are known to play a role in cellular mechanisms that control growth. The following descriptions of oncogene products are not meant to be complete, but to relate the above sections on mechanisms of oncogene activation to their role, or speculated role, in the transformation process. a. Growth Factors and Growth Factor Receptors The only retroviral oncogenes identified as genes that code for a peptide growth factor is the v-si_s oncogene of simian sacoma virus (SSV)(Devare et al., 1983) and the transforming gene of the Parodi-Irgenes feline sarcoma virus (Besmer et al., 1983), which both code for the B- chain of platelet-derived growth factor (PDGF). In vitro studies of SSV- transformed cells indicate that the mature v-sis gene product enters the cellular secretion pathway and interacts with the PDGF receptor to cause unrestricted growth in an autocrine fashion (Heldin et al., 1987). 17 Expression of the c-sis gene is not detectable in most normal cell lines or in most tissues (Eva et al., 1982), but has been found expressed in a number of human tumor cell lines (Heldin et al., 1987). In such cells expressing PDGF receptors, activation of c-sis gene expression may affect cells in a manner similar to SSV transformation, i.e. induce autocrine growth stimulation. A connection between oncogene products and growth factor receptors was established with the discovery that the v-aga oncogene of avian erythroblastosis virus, and the v-fms oncogene of the McDonough strain of feline sarcoma virus code for’ products homologous to the epidermal growth factor receptor (Ullrich et al., 1984; Xu et al., 1984: Lin et al., 1984) and colony stimulating factor-1 receptor (Rettenmeier' et al., 1985) respectively. In addition, the product of the gag oncogene, mentioned above, is thought to be the rat homolog of the human EGF receptor (Shector et al., 1984). The features common to the EGF and CSF-l growth factor receptors, as well as to the growth factor receptors for PDGF and insulin, is a glycosylated extracellular domain that binds ligand, a transmembrane domain consisting primarily of hydrophobic amino acid residues, and an intracelluar domain that exhibits intrinsic protein kinase activity. The tyrosine kinase activity of these receptors is dependent on the binding of the appropriate growth factor (Heldin et al., 1987), and appears to be an integral part of their function in stimulating cellular proliferation. However, there are many substrates and it has been difficult to find a molecular scheme to explain which tyrosine phosphorylations are important for such stimulation. In the case of the v-ag8 B oncogene, the region coding for the extracellular, ligand binding domain is deleted (Downward et al., 1984) and, as a consequence, its tyrosine kinase activity is 18 activated (Gilmore et al., 1985; Kris et al., 1985). Thus, alterations in the proportions of phoshorylated derivatives of EGF receptor-targeted intracellular proteins may act to cause cellular transformation by v-ag8 B. b. Intracellular Modulators of Cell Growth In addition to the growth factor receptor products of oncogenes, several other oncogene products that reside at the plasma membrane or in the cytoplasm code for protein tyrosine kinases, as well as serine/threonine kinases. Although the functions of these kinases are not known, it is thought that they control cell proliferation by modulating the activity or function of their various cytoplasmic and plasma membrane substrates. The oncogenic versions induce transformation through inappropriate or excessive phosphorylation of such substrates (Hunter, 1987). Therefore, it is not suprising that the gene that codes for protein kinase C, whose protein product is known to induce profound effects on cell growth and proliferation by substrate phosphorylation, can induce transformation of NIH3T3 cells if it is linked to strong viral transcriptional elements (Housey et al., 1988). The protein products of gas oncogenes, which are localized to the plasma membrane, appear to exert their profound effects on cell proliferation through a pathway involving protein kinase C and possibly other protein kinases. This will be discussed in a latter section. c. Nuclear Proteins Involved in Controlling Gene Expression The viral and cellular my; genes code for proteins that are located in the nucleus and have DNA-binding properties (Abrams et al. 1982). The 19 rapid increase of c-mys gene expression following mitogenic stimulation of resting cells (Leder et al. 1983) suggests that the c-mys protein plays a role in the cell cycle, possibly regulating entry into S-phase. As discussed above, the transforming properties of my; oncogenes are activated by unscheduled expression and/or overexpression. Thus, it is thought that such activation leads to the continuous supply of signals for entry into S-phase, and an inability of cells to cease dividing. Another nuclear oncogene product, that of the 188 oncogene, functions as a trans-acting transcriptional regulator (Angel et al., 1988). It was recently discovered that the cellular jag product forms a complex with the cellular product of another known nuclear oncogene, gas, and that this complex can interact with the regulatory sequences of many different genes (Rauscher et al, 1988). Like the cellular myc product, the cellular figs gene product is rapidly induced in various cell types in responce to mitogens (Mitchel et al., 1985; Muller et al., 1985; Treisman, 1985). Thus, interactions between the jag and figs gene products appear to be a part of a network of mechanisms that regulate the expression of genes involved in normal, and potentially abnormal, cellular proliferation in mammalian cells. 20 B. Biochemical properties of gas Using the NIH 3T3 transformation assay, gas oncogenes have been detected in approximately 10-15% of human tumors (Barbacid, 1987). The recent development and employment of more sensitive analytical methodologies for detection of gas genes indicate that a higher percentage of human tumors may contain gas oncogenes (Winter et al., 1985; 305 et al., 1986; 805 et al., 1987; Forrester et al., 1987). However, it is not possible to demonstrate a causative role for ,gas_ oncogenes in the development of human tumors simply by being able to detect their presence there. The use of carcinogen-induced animal tumor model systems provides one means of determining if there is a causal relationship between tumor induction and somatic mutation of gas genes. Several studies have shown that tumors which arise in animals following administration of chemical or other carcinogens contain activated gas genes, and that such activation is reproducible and carcinogen-specific (Balmain and Pragnell, 1983; Sukumar et al., 1983; Balmain et al., 1984; Zarble et al., 1985; Quintanilla et al., 1986; Roop et al., 1986). These in vivo studies provide strong evidence that gas genes play a role in mammalian tumorigenesis. 1. The gas Genes a. Structural Features of gas Genes Three functional gas genes have been identified in the human genome. These are designated H-gas, K-gas and N-gas (Shimizu et al., 1983). The H-gas gene, and the K-m gene were found to be homologous to the 21 transforming genes of Harvey murine sarcoma virus (Der et al., 1982, Parada et al., 1982; Santos et al., 1982) and Kirsten sarcoma virus (McCoy et al., 1983) respectively, hence their names. The N-gas gene was originally identified as a transforming gene in a human neuroblastoma cell line by the NIH 3T3 transformation assay (Hall et al., 1983). The three mammalian gas genes code for proteins that have a molecular weight of approximately 21 kilodaltons with 187 or 188 amino acid residues (Capon et al., 1983; Fasano et al., 1983; Reddy, 1983; Hall and Brown, 1984). The p21 coding sequences of each of the human gas genes is contained in 4 exons (K-gas also has an alternative 4th exon) with the first three exons being nearly identical, and each of the exon splice junctions for all three gas genes being identical (Capon et al., 1983; Fasano et al., 1983; Reddy, 1983; McGrath et al., 1983; Shimizu et al., 1983). Such sequence conservation suggests a common origin of r_as genes from one ancestral gene. Because variation in the size of the intervening sequences exists between H-, K- and N-gas genes, the genetic locus of these genes ranges in size from 4.5 kilobase pairs for H-gas to 50 kilobase pairs for K-gas (Fasano et al., 1983; McGrath et al., 1983; Shimizu et al.,1983; Capon et al., 1983). The promotor regions of each of the three functional mammalian gas genes lack a characteristic TATA or CAAT box (Breathnach and Chambon, 1981), but contain multiple transcriptioal start sites and copies of the hexanucleotide sequence CCGCCC or GGGCGG (GC box) (Hall and Brown, 1985; Ishii et al. 1985; McGrath et al., 1983). The GC box consensus sequence was originally found in the simian virus 40 early promoter (Byrne et al., 1983) and is characteristic of a number of cellular housekeeping genes (Ishii et al., 1986), 22 c. Mechanism of Activation of gas Genes Mammalian gas genes acquire the ability to transform cultured mammalian cells by single missense point mutations within their coding sequences (Tabin et al., 1982; Reddy et al., 1982; Taparowski et al., 1982). Sequence analysis of gas oncogenes derived from carcinogen induced tumors in animals, carcinogen transformed rodent cells, and from naturally occurring human tumors indicate that the mutations responsible for conferring transforming properties are located at codons 12 (Tabin et al., 1982; Reddy, et al., 1982; Taparowski et al., 1982), 13, (Bos et al., 1985), 59, (Dhar et al., 1982), and 61 (Yuasa et al., 1983). Missense mutations at several other positions, introduced by in vitro mutagenesis, also can confer transfoming properties to gas genes (Fasano et al., 1984; Walter et al., 1986; Sigal et al., 1986). Activation of gas genes by alterations affecting position 12 have been studied most thoroughly. Analysis of the transforming potential (focus formation in transfected Rat 1 cells) of 20 different H-gas genes, each encoding a different amino acid at this position, revealed that substitution of the normal amino acid (glycine) with any one of the other 20 amino acids, except proline, resulted in an activated p21 (Seeburg et al., 1984). There is some evidence that overexpression of normal gas genes, can play a role in cell transformation in vitro and tumor development in vivo. For example, it has been shown that normal gas genes can cause malignant transformation of NIH 3T3 cells when linked to a strong promotor (Chang et al., 1982), or when multiple copies are incorporated into the genome 23 of these cells (Pulciani et al, 1985). Transformation in each of these cases correlated with a 30-100-fold increase in the level of g_s expression over that observed in non-transfected NIH 3T3 cells (Chang et al., 1982; Pulciani et al., 1985). Amplification and overexpression of normal gas genes, usually K-gas, has been observed in a number of human tumors (Kahn et al., 1987), but the levels of expression were much below that which was capable of causing transformation of NIH 3T3 cells. Furthermore, there has been no consistent results from studies examining gas expression in tumor tissues that would implicate a role for overexpression of normal gas alleles in tumorigenesis (Barbacid, 1987). 2. The gas Proteins 3. Structure of gas Proteins Several different functional domains of gas p21 proteins have been identified (Figure 1). These domains were identified using primarily three approaches: comparing the deduced amino acid sequence of p215 that exhibit transforming properties and those that do not (Taparowski et al. 1982; Dhar et al., 1982; Seeburg, et al., 1984; Fasano et al., 84; Walter et al., 1984; Sigal et al., 1986); comparing the biochemical properties of mutant proteins expressed in as 2911 with their biological properties in cells microinjected with the mutant protein or in cells transfected with the mutated gene (Willumsen et al., 1985; Willumsen et al., 1986; Sigal et al., 1986; Der et al., 1986); and determining the biochemical and biological properties of gas p21 following reaction with antibodies with known p21 epitopes (Clark et al., 1985). The binding domain of the Y13- 259 monclonal antibody (Furth et al., 1981), which has been utilized extensively for characterization of gas proteins, is located in a highly 24 no I I I I I I “am" ' to muons ODflGfl' "”““““'M”P Efloetor vnpzn umuuyuumn 5:3 Membrane “WI MW nglon Figure 1. Schematic representation of the structural and functional domains within mammalian p21 gas proteins. Printed with permission from Ann. Rev. Biochem. Inc. 25 conserved, but functionally expendible region (Willumsen et al., 1986; Sigal et al., 1986; Lacal et al., 1986). The only region of gas proteins that shows sequence diversion (hypervariable region) is near the carboxy terminus, suggesting that some as yet unknown feature of this region will provide an explanation as to why three so functionally similar gas genes are expressed in mammalian cells. b. GTP Binding and Hydrolysing Activities The known biochemical activities associated with gas p21 proteins include GDP and GTP binding (Scolnick et al, 1979; Shih et al. 1980) and GTPase activities (McGrath et al., 1984; Gibbs et al., 1984; Sweet et al., 1984). In addition, gas proteins possesing threonine at amino acid position 59 (v-H-gas p21) exhibit autophosphorylation at that residue (Shih et al., 1982). In all g_s p21’s, palmitic acid is post- translationally attached to a cysteine residue near the carboxyl terminus which facilitates their association with the inner surface of the plasma membrane (Willingham et al., 1980; Shih et al., 1982; Sefton et al., 1982; Willumsen et al., 1984). A biochemical marker that seems to correlate well with mutations that activate the transforming properties of gas oncogenes is decreased GTPase activity (McGrath et al., 1984; Gibbs et al., 1984: Sweet et al., 1984). The biochemical properties of ES p21 proteins closely resemble those of the G proteins involved in transmembrane signaling systems (Gilman, 1984). Such G proteins cycle between active GTP-bound forms, and inactive GDP-bound forms, and are involved in the stimulation or 26 inhibition of adenyl cyclase (Gilman, 1984). The observation that many oncogenic mutants of p21 are reduced in GTPase activity led to the proposal that transformation by p21 is the result of abnormally high levels of p21 in the active GTP-bound state (Gibbs et al., 1984; McGrath et al.,1984; Sweet et al., 1984). However, analysis of GTPase activities associated with a large number of p21 mutants failed to reveal a clear quantitative relationship between GTPase activity and transforming potential (Trahey et al. 1987; Colby et al., 1986; Der et al., 1986). Furthermore, no such association between mammalian gas p21 proteins and adenylate cyclase has been observed (Beckner et al., 1985). c. Interaction with a Putative Effector Protein The putative gas p21 effector domain, the region or regions that interact with molecules either upstream or downstream in the gas pathway, was ascribed to a region of the NHZ-terminal end of gas proteins (Figure 1) by analyzing deletion mutants (Sigal et al., 1986). Mutations in the region that includes amino acids 35-40 reduce the biological effects of gas proteins in NIH 3T3 cells, but do not disrupt the in vitro biochemical properties of 8. coli expressed gas proteins (Sigal et al., 1986). Recently, it was discovered that a protein, designated GAP (GTPase activating protein) interacts with the gas p21 (Trahey et al., 1987), and that such interaction occurs at the putative effector domain (Cales et al., 1988). The GAP protein is a cytosolic protein that stimulates the GTPase of normal gas p21 by loo-fold with no apparent effect on the GTPase of oncogenic gas p21 (Trahey et al., 1987). The studies by Cales et al. (1988) suggest that association of GAP with the normal gas p21-GTP bound 27 complex leads to a conformational change in the protein that increases its intrinsic GTPase activity, and that due to the mutations in the oncogenic forms of ras p21, such conformational changes are prevented from occurring, resulting in a reduced intrinsic GTPase activity. Formal proof of a role for GAP as the target or a regulator of gas p21 has yet to be obtained. d. Involvement in the Control of Cell Proliferation The normal physiological role of the gas proteins remains unknown. Studies on the expression of gas genes in mammalian cells indicate that they are expressed at low levels, in most, if not all cell lineages (Chesa et al., 1987; Furth et al., 1987). Such data imply that the physiological role of ,gas, proteins is shared by most cell types. Because the biochemical properties of {as genes are similar to those of proteins involved in signal transduction, and because expression of ras oncogenes in various rodent cells lead to the abnormal growth characteristics exhibited by transformed cells, gas genes are thought to participate in biochemical pathways essential to normal cellular proliferation. It has been known for a long time that various growth factors cause changes in inositol lipid metabolism (Durell et al., 1969). More recently, it has been shown that biochemical activities correlated with the action of GTP binding/hydrolyzing proteins are involved in pathways linked to phosphotidylinositol turnover (Brown et al., 1984). Intermediates in this pathway, particularly diacylglycerol (DAG) and inositol 1,4,5-triphospate (1P3), can act as second messengers to mediate cellular responses initiated by the binding of growth factors, as well as 28 neurotransmitters and hormones to their cell surface receptors (Berridge, 1984). In as transformed NIH3T3 cells the steady state levels of phoshotidylinositol 4,5-diphosphate (PIPZ) and DAG are elevated compared to their untransformed counterparts (Fleischman et al., 1986), and it has been suggested that gas p21 may act to affect such levels by activation of a specific enzyme, phospholipase C (Fleischman et al., 1986), which catalyses the breakdown of PIP2 to IP3 and DAG. IP3 itself stimulates release of calcium into the cytosol from intracellular stores (Berridge, 1984) and DAG activates a calcium- and phospholipid-dependent kinase (protein kinase C) (Nishizuka, 1986) and acts as a substrate for arachidonic acid release during prostaglandin synthesis. Both the DAG- activated protein kinase C and the IP3-stimulated Ca“ release cause a plasma membrane NaI-H’ exchanger, which elevates cytoplasmic Na+ and raises the pH (Streb et al., 1983). Recent evidence suggests that generation of these signals through stimulation of phosphinositol turnover is a critical event in growth factor stimulation of cell proliferation (Berridge et al., 1984). Although there is little direct data supporting a role for gas genes in this pathway, it has been speculated that the loss of GTPase activity in oncogenic gas proteins causes these proteins to activate phospholipase C and the concomitant production of 1P3 and DAG in an uncontrolled way, independent of signals supplied by growth factor binding (Berridge, 1984; Fleischman et al., 1986). 3. Biological Properties Associated with Expression or gas Oncogenes in Mammalian Cells a. Biological Properties Induced in Rodent Fibroblasts 29 Studies by several groups (Land et al., 1983; Ruley et al., 1983; Newbold and Overell, 1983; Thomassen et al., 1985) have indicated that transfection of gas oncogenes into early passage rodent fibroblasts can cause such cells to exhibit some properties of tumor cells, such as a transformed morphology and anchorage independence, but not the infinite lifespan phenotype or tumorigenicity. Malignant transformation of finite lifespan rodent fibroblasts can be attained if a ,gas oncogene is transfected along with any one of a number of different genes, such as v- myc, c-myc or polyoma large T (Land et al., 1983), adenovirus ElA (Ruley, 1983), N-myc (Schwab et al., 1985), or p53 (Parada et al., 1985), that are thought to be capable of causing the infinite lifespan phenotype in rodent cells in culture. Malignant transformation can also result following transfection of gas oncogenes into rodent fibroblasts that already exhibit an infinite lifespan in culture (Land at al., 1983; Newbold and Overell, 1983). Together, these results suggest that the infinite lifespan phenotype is a prerequisite for’ malignant transformation of rodent fibroblasts induced by expression of a transfected ras oncogene. However, it has been reported that transfection of a ga_s oncogene alone is sufficient for the malignant transformation of finite lifespan mouse, rat, or Chinese hamster fibroblasts (Pozatti et al., 1986; Spandidos and Wilkie, 1984; Land et al., 1986; Spandidos, 1985), and it appears that a high level of gas oncogene expression was necessary for such transformation. Such malignant transformation has not been achieved following transfection of gas oncogenes into Syrian hamster fibroblasts (Thomassen et al., 1985, Newbold and Overell, 1983), which have a more stable karyotype and a lower incidence of spontaneous transformation into cells having an infinite 30 lifespan than do mouse, rat or Chinese hamster fibroblasts (Newbold and Overell, 1983; Thomassen et al., 1985). This latter result has led to speculation that the malignant transformation reported to occur following transfection of a gas oncogene into finite lifespan mouse, rat and Chinese hampster fibroblasts was caused by the increased propensity of such fibroblasts to aquire an infinite lifespan (Thomassen et al., 1985). Therefore, the hypothesis that the acquisition of an infinite lifespan in culture is a prerequisite for the malignant transformation of rodent fibroblasts is still widely accepted. b. Biological Consequences of Expression of gas Oncogenes in Transgenic Mice An important limitation of experiments involving the use of cultured cells for evaluating the action of oncogenes, is that they do not accurately reflect transformation as it occurs in vivo. Such experiments cannot assess the influence of various factors, such as blood supply, humoral and cell-mediated Igrowth controls, and physical barriers to growth, that are thought to be crucial to the development of tumors in vivo (Stewart et al., 1984). In contrast, the use of transgenic mice, i.e., mice containing exogenous sequences introduced into their germ lines, permits the evaluation of the action of oncogenes in the context of the living organism. Three studies utilized transgenic mice to investigate the effects of gas oncogenes (Quaife et al., 1987; Andres et al., 1987; Sinn et al., 1987). In each of these studies an H-gas oncogene, linked to transcriptional regulatory elements that direct expression in a tissue- 31 specific manner, was introduced into the germ line of mice (Quaife et al., 1987; Andres et al., 1987; Sinn et al., 1987). Expression of an H-gas oncogene under the control of the rat elastase regulatory elements which direct transcription only in the pancreas (MacDonald et al., 1987), resulted in the neoplastic proliferation of all, or most of the pancreatic acinar cells (Quaife et al., 1987). However, such neoplastic proliferation does not necessarily represent a true malignant tumor, but more likely represents a benign, but lethal neoplasm. In contrast to these results, expression of an H-gas oncogene under the control of either a mouse mammary virus tumor (MMTV) promoter (Sinn et al., 1987), or under the control of the mouse whey acidic protein promotor (Andres et al., 1987), each of which directs transcription primarily in mammary epithelium, resulted primarily in the development of mammary tumors (adenocarcinomas) that were apparently of clonal origin and occurred only after a relatively long latent period (usually 6-12 months) (Sinn et al., 1987; Andres et al., 1987). Such results suggest that the cells of certain tissues are capable of responding directy to the transforming effects of gas oncogenes by exhibiting benign neoplastic growth and that other cells require additional changes at the cellular level before neoplastic proliferation or malignant tumor development occurs. In experiments by Sinn et al. (1987), mice matings were performed between transgenic mice that harbored an H-gas oncogene linked to the MMTV promoter, and mice that harbored a cellular mys gene linked to the same promoter (Stewart et al., 1984), to examine the combined effects of these genes in target cells of progeny animals. Coexpression of these genes resulted in'a change in the spectrum of cell types that developed tumors 32 and a decrease in the time of tumor onset compared to that observed in animals expressing the gas or my; transgenes separately (Sinn et al., 1987). Nevertheless, the tumors formed appeared to be of clonal origin, and tumor incidence in these animals increased in an age dependent manner (Sinn et al., 1987), suggesting that events additional to co-expression of the gas and mys transgenes were required for tumor development by the targeted cells. c. Use of Human Cells to Study the Biological Effects of gas Oncogene expression For obvious reasons humans cannot be experimentally manipulated in the same manner as mice or other animals. Therefore, in the case of the human species, cell culture transformation systems take on special relevance for the study of cancer. However, no reproducible method has been developed to bring about the malignant transformation of human cells in culture using carcinogenic agents or transfection oncogenes (DiPaolo, 1984; McCormick and Maher, 1988). The failure of human fibroblasts to transform as readily as the rodent cell lines to date prevents a direct comparison of the tranformation process in human and rodent cells, and raises the question of whether the results achieved using rodent cells can be applied to cancer development in humans. Such dissimilarities between rodent and human cell transformation systems may reflect a fundamental biological difference between the two species, or may merely relect a significant difference in the frequency of acquiring some required phenotype, such as infinite lifespan (McCormick and Maher, 1988). There have been several studies that have attempted to identify the 33 biological consequences of expression of transfected gas oncogenes in human fibroblasts. These studies are discussed in Chapters 11 and III of the thesis. REFERENCES Abrams, U., Rohrschneider, L. and Eisenman, R.N. (1982) Nuclear location of the putative transforming protein of avian nwelocytomatosis virus. Cell 88, 427-439. Anderson, P., Goldfarb, N.P. and Weinberg, R.A. (1979) A defined subgenomic fragment of in vitro synthesized Moloney sarcoma virus DNA can induce cell transformation upon transfection. Cell 18, 63-75. Andres, A.C., Schonenberger, C.A., Groner, B., Hennighausen, L., LeMeur, M. and Gerlinger, P. (1987) Ha-gas oncogene expression directed by a milk protein gene promoter: Tissue specificity, hormonal regulation, and tumor induction in transgenic mice. Proc. Natl. Acad. Sci. 88, 1299-1303. Angel, P., Allegretto, E.A., Okino, S.T., Hattori, K., Boyle, W.J., Hunter, T. and Karin, M. (1988) Oncogene jgg_encodes a sequence-specific trans-activator similar to AP-l. Nature 888, 166-171. Balmain, A., Ramsden, M., Bowden, G.T. and Smith, J. (1984) Activation of the mouse cellular Harvey-gas gene in chemically induced benign skin papillomas. Nature 881, 658-660. Balmain, A. and Pragnell, 1.8. (1983) Mouse skin carcinogenesis in duced in vivo by chemical carcinogens have a transforming Harvey gas oncogene. Nature, 888, 72-74. Baltimore, D. (1970) Viral RNA-depencent DNA ploymerase. Nature 888, 1209-1215. Barbacid, M. (1987) gas Genes Ann. Rev. Biochem. 88, 779-827. Bargeman, C.I., Humg, M.C. and Weinberg, R.A. (1986) Multiple independent activations of the gag oncogene by a point mutation altering the transmembrane domain of p185. Cell 88, 649-657. Barrett, J.C. and Ts’o, P.0.P. (1978) Relationship between transformation and somatic mutation in human and Chinese hamster cells. Proc. Natl. Acad. Sci. USA 18, 3297-3302. Beckner, S.K., Hattori, S. and Shih, T.Y. (1985) The m oncogene product p21 is not a regulatory component of adenylate cyclase. Nature 811, 71-72. Berenblum, I. and Shubik, P.A. (1947) A new, quanitative, approach to the study of the stages of chemical carcinogenesis in the mouse’s skin. Br. J. Cancer 1, 383-391. Berenblum, I. (1940) The mechanism of carcinogenesis: A study of the significance of carcinogenic action and related phenomenon. Cancer Res. 1, 45-53. 34 35 Berridge, M.J. (1984) Inositol triphosphate and diacylglycerol as second messengers. Biochem. J. 888, 345-360. Besmer, P., Snyder, H.W., Murphy, Hardy, W.D. and Parodi, A.J. (1983) The Parodi-Irgens feline sarcoma virus and simian sarcoma virus have homologous oncogenes, but in different contexts of the viral genome. J. Virology 88, 606-612. Bishop, J.M. (1987) The molecular genetics of cancer. Science 888, 305- 311. 805, J.L., Tokoz, 0., Marshall, C.J., Ver-de Vries, M., Veeneman, G.H., van der Eb, A.J., van Boom, J.H., Janssen, W.G. and Steenvoorden. (1985) Amino acid substitution at codon 13 of the N-gas oncogene in acute myeloid leukemia. Nature 818, 726-730. 805, J.L., Fearon, E.R., Hamilton, S.R., Verlaan-deVries, M., van Boom, J.H., van der EB, A.J. and Vogelstein, B. (1987) Prevalence of gas gene mutations in human colorectal cancers. Nature 881, 293-298. Bos, J.L., Verlaan-de Vries, M., Jansen, A.M., Veeneman, G.M., van Boom, J.L and van der Eb, A.J. (1986) A human gastric carcinoma contains a single point mutated and amplified normal allele of the Ki-g_s oncogene. Nucl. Acids Res. 18, 1209-1217. Breathnach, R and Chambon, P. (1981) Organization and expression of eucaryotic split genes coding for proteins. Ann. Rev. Biochem. 88, 349- 383. Brodeur, G.M., Seeger, M., Schwab, M., Varmus, H.E. and Bishop, J.M. (1984) .Amplification of N-myc in untreated neuroblastomas correlates with advanced disease stage. Science 888, 1121-1124. Brown, R.D., Berger, K.D. and Taylor, P.J. (1984) p-Adrenergic receptor activation mobilizes cellular Ca"in a muscle cell line. J. Biol. Chem. 888, 7554-7562. Byrne, B.J., Davis, M.S., Yamaguchi, J., Bergsma, D.J. and Subramanian, K.N. (1983) Definition of the simian virus 40 early promoter region and demonstration of a host range bias in the enhancement effect of the simian virus 40 72-base-pair repeat. Proc. Natl. Acad. Sci. 88, 721-725. Cairns, J. (1981) The origins of human cancer. Nature 888, 353-357. Cales, C., Hancock, J.F., Marshall, C.J. and Hall, A. (1988) The cytoplasmic protein GAP is implicated as the target for regulation by the gas gene product. Nature 888, 548-550. Capon, D.J., Seeburg, P.H., McGrath, J.P., Hayflick, J.S, Edman, U., Levinson, A.D. and Goeddel, D.V. (1983) Activation of Ki-gas-Z gene in human colon and lung carcinoma by two different point mutations. Nature 881, 507-512. 36 Chan, L.C., Karhi, K.K., Rayter, S.I., Heisterkamp, N., Eriani, S., Powles, R., Lawler, S.D., Groffen, J., Foulkes, J.G., Greaves, M.F. and Wiedemann, L.M. (1987) A novel a81 protein expressed in Philadelphia chromosome positive acute lymphoblstic leukemia. Nature 888, 635-637. Chang, E.H., Furth, M.E, Scholnick, E. and Lowy, D. (1982) Tumorigenic transformation of mammalian cells induced by a normal human gene homologous to the oncogene of Harvey murine sarcoma virus. Nature 881, 479-484. Chesa, P.G., Rettig, W.J., Melamed, M.R., Old, L.J. and Niman, H.L. (1987) Expression of p21 gas in normal and malignant tissues: Lack of association with proliferation and malignancy. Proc. Natl. Acad. Sci. 88, 3234-3238. Cohen, J.B., Unger, T., Rechavi, G., Canaani, E. and Givol, D. (1983) Rearrangement of the oncogene c-mas in mouse melanoma N51 and hybridomas. Nature 888, 797-799. Colby, W.W., Hayflick, J.S., Clark, 5.6. and Levinson, A.D. (1986) Biochemical characterization of polypeptides encoded by mutated human Ha- gas 1 genes. Mol. Cell. Biol. 8, 730-734. Cooper, C.S., Park, M., Blair, D.G., Tainsky, M.A., Huebner, K., Croce, C.M. and Vande Woude. (1984) Molecular cloning of a new transforming gene from a chemically transformed human cell line. Nature 811, 29-33. Croce, C.M., Shander, M., Martinis, J., Cicurel, L., D’Anacona, G.C., Dolby, T.W., and Koprowski, H. (1979) Chromosomal location of the genes for human immunoglobulin heavy chains. Proc. Natl. Acad. Sci. 18, 3416- 3419. Dalla-Favera, R., Wong-Staal, F. and Gallo, R.C. (1982) 888 gene amplification in promyelocytic leukemia cell lines HL-60 and primary leukemia cells of the same patient. Nature 888, 61-63. Davis, R.L., Konopka, J.B. and Witte, 0. (1985) Activation of the c-a81 by viral transduction or chromosomal translocation generates altered c- a81 proteins with similar in vitro kinase properties. Mol. Cell. Biol. 8, 204-213. de Klein, A., Geurts van Kessel, A., Grosveld, G., Bartram, C.R., Hagemeijer, A., Bootsma, D., Spurr, N.K., Heisterkamp, N., Groffen, J. and Stephenson, J.R. (1982) A cellular oncogene is translocated to the philadelphia chromosome in chronic myelogenous leukemia. Nature 380, 765- 769. Der, C.J., Krontirus, T.G. and Cooper, G.M. (1982) Transforming genes of human bladder and lung carcinoma cell lines are homologous to the gas genes of Harvey and Kirsten sarcoma virus. Proc. natl. Acad. Sci. 18, 3637-3640. 37 Der, C.J., Finkel, T. and Cooper, G.M. (1986) Biological and biochemical properties of human gas" genes mutated at codon 61. Cell 4_4_, 167-176. Devare, S.G., Reddy, E.P., Robbins, K.C., Andersen, P.R., Tronick, S.R. and Aaronson, S.A. (1983) Nucleotid sequence of the simian sarcoma virus genome: Demonstration that its acquired cellular sequences encode the transforming gene product p28 sis. Proc. Natl. Acad. Sci. 88, 731-735. Dhar, R., Ellis, R.W., Shih, T.Y., Oroszlan, S. and Shapiro, B. (1982) Nucleotide sequence of the oncogene encoding the p21 transforming protein of kirsten murine sarcoma virus. Science 811, 934-936. DiPaolo, J.A. (1984) Relative difficulties in transforming human and animal cells in vitro. J. Natl. Cancer Inst. 18, 3-8. Doll, R. (1978) An epidemiological perspective of the biology of cancer. Cancer Res. 88, 3573-3583. Downward, J., Yarden, Y., Mayes, E., Scrace, G., Totty, N., Stockwell, P., Ullrich, A., Schlessinger, J. and Waterfield, M.D. (1984) Close similarity of epidermal growth factor receptor and v-erbB oncogene protein sequences. Nature 881, 521-528. Downward, J., Parker, P. and Waterfield, M.D. (1984) Autophosphorylation sites on the epidermal growth factor receptor. Nature 811, 483-485. Durel, J., Garland, J.T. and Friedel, R.0. (1969) Acetylcholine action: Biochemical aspects. Science 188, 862-866. Erickson, J., Martinis, J. and Croce, C.M. (1982) Assignment of the genes for immunoglobulin chains to chromosome 22. Nature 888, 173-175. Eva, A., Robbins, K.C., Andersen, P.R., Srinivasan, A., Tronick, S.R., Reddy, E.P., ellmore, N.W., Galen, A.T., Lautenberger, J.A., Papas, T.S., Westin, E.H., Wong-Staal, F., Gallo, R.C. and Aaronson, S.Au (1982) Cellular genes analogous to retroviral gas genes are transcribed in human tumor cells. Nature 888, 116-119. Eva, A and Aaronson, S.A. (1985) Isolation of a new human oncogene from a diffuse B-cell lympoma. Nature 818, 273-275. Fasano, 0., Taparowski, E, Fiddes, J., Wigler, M. and Goldfarb, M. (1983) Sequence and structure of the coding region of the human H-gas gene from T24 bladder carcinoma cell. J. Mol. Appl. Genet. 8, 173-180. Fasano, 0., Aldrich, T., Tamanoi, F., Taparowski, E., Furth, M. and Wigler, M. (1984) Analysis of the transforming potential of the human H-gas gene by random mutagenesis. Proc. Natl. Acad. Sci. USA 81, 4008- 412. 38 Feramisco, J.R., Clark, R., Wong, G., Arnheim, N., Milley, R. and McCormick, F. (1985) Transient reversion of ,gas oncogen-induced transformation by antibodies specific for amino acid 12 of gas protein. Nature 818, 639-642. Fleischman, L.F., Chahwala, 5.3., and Cantley, L. (1986) ,gas transformed cells: Altered levels of phosphotidylinositol-4,5-bisphosphate and catabolites. Science 881, 407-410. Forrester, K., Almoguera, C., Han, K., Grizzle, H.E. and Perucho, M. (1987) Detection of high incidence of K-ras oncogenes during human colon tumorigenesis. Nature 881, 298-303. Fung, Y.K., Lewis, W.G., Kung, H.J. and Crittenden, L.B. (1983) Activation of the cellular oncogene c-er88 by LTR insertion: Molecular basis for induction of erythroblastosis by avian leukosis virus. Cell 88, 357-368. Furth, M.E., Aldrich, T.H. and Cordon-Cardo, C. (1987) Expression of gas proto-oncogene proteins in hormal human tissues. Oncogene 1, 47-58. Fusco, A., Grieco, M., Santoro, M., Berlingieri, M.T., Pilotti, S., Pierotte, M.A., Della Porta, G. and Vecchio, G. (1987) A new oncogene in human thyroid papillary carcinomas and their lymph-nodal metastases. Nature 888, 170-173. Gattoni-Celli, S., Hsiao, W.L. and Weinstein. (1983) Rearranged c-mos locus in a MOPC 21 murine myeloma cell line and its persistence in hybridomas. Nature 888, 795-797. Gibbs, J.B., Sigal, I.S., Poe, M. and Scolnick, E.H. (1984) IntrinSic GTPase activity distinguishes normal and oncogenic p21 gas molecules. Proc. Natl. Acad. Sci. USA 81, 5704-5708. Gilman, A.G. (1984) G proteins and dual control of adenylate cyclase. Cell 88, 577-579. Gilmore, T., DeClue, J.E. and Martin, G.S. (1985) Protein phosphorylation at tyrosine is induced by the v-arbB Igene product in vivo and in vitro. Cell 88, 609-618. Goldfarb, M., Shimizu, K., Perucho, M. and Wigler, M. (1982) Isolation and preliminary characterization of a human transforming gene from T24 bladder carcinoma cells. Nature 828, 404-409. Graham, F.L. and Van der Eb, A. (1973) A new technique for the assay of infectivity of human acenovirus 5 DNA. Virology 88, 456-461. Hall, A. and Brown R. (1985) Human N-m: c-DNA cloning and gene structure. Nucleic Acids Res. 18, 5255-5268. Hall, A., Marshall, C., Spurr, N. and Weiss, R. (1983) Identification of the transforming gene in two human sarcoma cewll lines as a new member of the gas family located on chromosome 1. Nature 888, 396-400. 39 Haseltine, W.A. (1983) Ultraviolet light repair and mutagenesis revisited. cell 88, 13. Hayward, W.G., Neel, B.E. and Astrin, S.M. (1981) Activation of a cellular gas gene py promoter insertion in ALV-induced lymphoid leukosis. Nature 888, 475-480. Heldin, C.H., Betsholtz, C., Claesson-Welsh, L. and Westermark, B. (1987) Subversion of growth regulatory pathways in malignant transformation. Biochim. Biophys. Acta. 881, 219-244. Housey, G.M., Johnson, M.D., Hsiao, W.L.W., O’Brian, C.A., Murphy, J.P., Kirshmeier, P, and Weinstein, 1.8. (1988) Overproduction of protein kinase C causes disordered growth production in rest fibroblasts. Cell 88, 343-354. Huebner, R.J. and Todaro, G.J. (1969) Oncogenes of RNA tumor viruses as determinants in cancer. Proc. Natl. Acad. Sci. USA 88, 1087-1092. Hunter, T. (1987) A thousand and one protein kinases. Cell 88, 823-829. Ishii, S., Kadonaga, J.T., Tjian, R., Brady, J.N., Merlino, G.T., and Pastan, I. (1986) Binding of the Spl transcription factor by the human Harvey gas 1 proto-oncogene promoter. Science 888, 1410-1413. Ishii, S., Merlino, G.T. and Pastan, I. (1985) Promoter region of the human Harvey gas proto-oncogene: Similarity to the EGF receptor proto- oncogene promoter. Science 888, 1378-1382. Kelly, K., Cochran, B.H., Stiles, C.D. and Leder, P. (1983) Cell- specific regulation of the c-myc gene by lymphocyte mitogens and platelet- derived growth factor. Cell 88, 603-610. Kier, L.D., Yamasaki, E. and Ames, B.N. (1974) Detection of mutagenic activity in cigarette smoke condensates. Proc. Natl. Acad. Sci. 11, 4159- 4163. Kris, R.M., Lax, 1., Gullick, W., Waterfield, M.D., Ullrich, A., Fridkin, M. and Schlessinger, J. (1985) Antibodies against a synthetic peptide as a probe for the kinase activity of the avian EGF receptor and the v- egb8 protein. Cell 58, 619-625. Kurzrock, R., Shtalrid, M., Romero, P., Kloetzer; W.S., Talpas, M., Trujillo, J.M., Blick, M., Beran, M. and Gutterman, J.U. (1987) A novel protein product in Philadelphia-positive acute lymphoblastic leukemia. Nature 888, 631-633. Land, H., Parada, L.F. and Weinberg., R.A. (1983) Tumorigenic conversion of primary embryo fibroblasts requires at least two cooperating genes. Nature 885, 596-602. Land, H., Chen, A.C., Morgenstern, J.P., Parada, L.F. and weinberg, R.A. (1986) Behaviour of my; and gas oncogenes in transformation of rat embryo fibroblasts. Mol. Cell. Biol. 6, 1917-1925. 4o Leder, P, Battey, J., Lenoir, G., Moulding, C., Murphy, W., Potter, H., Steward, T. and Taub, R. (1983) Translocations among antibody genes in human cancer. Science 888, 765-771. Little, C.D., Nau, M.M., Carney, D., Gazdar, A.F. and Minna, J.D. (1983) Amplification and expression of the c-mys oncogene in human lung cancer cell lines. Nature 888, 194-196. Maher, V.M., Quellette, L.M., Curren, R.D. and McCormick, J.J. (1976) Frequency of ultraviolet light-induced mutations is higher in xeroderma pigmentosum variant cells than in normal cell. Nature 881, 593-595. Malcolm, S., Barton, P., Murphy, C. Fergusson-Smith, M.A., Bentley, D.L. and Rabbitts, T.H. (1982) Localization of human immunoglobulin light chain variable region genes to the short arm of chromosome 2 by in situ hybridization. Proc. Natl. Acad. Sci. USA 18, 4957-4961. Marcu, K.B., Harris, L.J., Stanton, L.W., Erickson, J., Watt, R. and Croce, C.M. (1983) Transcriptionally active c-mys oncogene is contained within NIARD, a DNA sequence associated with chromosome translocations in B-cell neoplasia. Proc. Natl. Acad. Sci. 88, 519—523. Martin-Zanca, D., Hughes, S.H. and Barbacid, M. (1986) A human oncogen foprmed by the fusion of truncated tropomyosin and protein tyrosine kinase sequences. Nature 818, 743-748. McCormick, J.J. and Maher, V.M. (1988) Towards an understanding of the malignant transformation of diploid human fibroblasts. Mutation Res. in press. McCoy, M., Toole, J.J., Cunnigham, J.M., Chang, E.H., Lowy, D.R. and Weinberg, R.A. (1983) Characterization of a colon/lung carcinoma oncogene. Nature 888, 79-81. McGrath, J.P., Capon, D.J., Smith, D.H., Chen, E.Y., Seeburg, P.H., Goeddel. D.V. and Levinson, A.D. (1983) Structure and organization of the human Ki-gas proto-oncogene and a related ppseudogene. Nature 888, 501-506. McGrath, J.P., Capon, D.J., Goeddel, D.V. and Levinson, A.D. (1984) Comparative biochemical properties of normal and activated human gas p21. Nature 818, 644-649. Mitchell, R.L, Zodas, L., Schreiber, R.D. and Verma,I.M. (1985) Rapid induction of the expression of proto-oncogene gas during human monocytic differentiation. Cell 88, 209-217. Moolgavkar, S.H. and Knudson, A.G. (1981) Mutation and cancer: A model for human carcinogenesis. J. Natl. Cancer Inst. 88, 1037-1052. Muller, R. and Wagner., E.F. (1984) Differentiation of F9 teratocarcinoma stem cells after transfer of c-fas_oncogenes. Nature 811, 438-442. 41 Neel, B.G., Hayward, W.S., Robinson, H.L., Fang, J. and Astrin, S.M. (1981) Cell 88, 323-334. Neiman, P.E. (1972) Rous sarcoma nucleotide sequences in cellular DNA: measurement by RNA-DNA hybridization. Science 118, 750-752. Newbold, R.F and Overell, R.W. (1983) Fibroblast immortality is a prerequisite for transformation by EJ c-Ha-gas oncogene. Nature 888, 648- 641. Nishizuka, Y. (1986) Studies and perspectives of protein kinase C. Science 888, 305-312. Nowell, P.C. and Hungerford, D.A. (1960) A minute chromosome in chronic granulocytic leukemia. Science 188, 1497-1500. Padhy, L.C., Shih, C., Cowing, D., Finkelstein, R. and Weinberg, R.A. (1982) Identification of a phosphoprotein specifically induced by the transfoming DNA of rat neuroblastoms. Cell 88, 865-871. Padua, R.A., Barrass, N. and Currie, G.A. (1984) A novel transforming gene in a malignant melanoma cell line. Nature 811, 671-673. Parada, L.F., Land, H., Weinberg, R.A., Wolf, 0. and Rotter, V. (1985) Cooperation between gene encoding p53 tumor antigen and gas in cellular transformation. Nature 818, 649-651. Parada, L., Tabin, C., Shih, C. and Weinberg, R.A. (1982) Human EJ bladder carcinoma oncogene is a homologue of the Harvey sarcoma virus gas gene. Nature 881, 474-478. Payne, G.S., Bishop, J.M. and Varmus, H.E. (1982) Multiple arrangements of viral DNA and an activated host oncogene in bursal lymphomas. Nature 888, 209-214. Pozzatti, R., Muschel, R., Williams, J., Padmanabhan, R., Howard, B., Liotta, L. and Khoury, G. (1986) Primary rat embryo cells transformed by one of two oncogenes show different metastatic potentials. Science 888, 223-227. Pulciani, S., Santos, E., Long, L., Sorrentino, V. and Barbacid, M. (1985) gas gene amplification and malignant transformation. Mol. Cell. Biol. 8, 2836-2841. Quaife, C.J., Pinkert, C.A., Ornitz, D.M., Palmiter, R.D. and Brinster, R.L. (1987) Pancreatic neoplasia induced by gas expression in acinar cells of transgenic mice. Cell 88, 1023-1034. Quintanilla, M., Brown, K., Ramsden, M. and Balmain, A. (1986) Carcinogen-specific.mutation and amplification of'Ha-gas_during mouse skin carcinogenesis. Nature 888, 78-80. Rauscher, F., Cohen, D.R., Curren, T., Bos, T.J., Vogt, P.K., Bohmann, D., Tjian, R. and Franza, R. (1988) Fos-associated protein p39 is the product of the jug proto-oncogene. Science 888, 1010-1016. 42 Reddy, P., Reynold, R.K., Santos, E. and Barbacid, M.A. (1982). A point mutation is responsible for the acquisition of transforming properties by the T24 H-gas human bladder carcinoma oncogene. Nature 888, 149-152. Reddy, P. (1983) Nucleotide sequence analysis of the T24 human bladder carcinoma oncogene. Science 881, 1061-1064. Rettenmeir, C.W., Sacca, R., Roussel, M.F., Look, A.T. and Stanley, E.R. (1985) The c-fms proto-oncogene product is related to the receptor for the mononuclear phagocyte growth factor , CSF-l. Cell 81, 665-676. Roop, D.R., Lowy, D.R., Tambourin, P.E., Strickland, J., Harper, J.R., Balasckhak, M., Spangler, E.F. and Yuspa, S.H. (1986) An activated Harvey gas oncogene produces benign tumors on mouse epidermal tissues. Nature 888, 822-824. Ruley, H.E. (1983) Adenovirus early region 1A enables viral and cellular transforming genes to transform primary cells in culture. Nature 888, 602-606. Sandberg, A.A. (1980) The chromosomes in human cancer and leukemia. Elsevier, New York. Santos, E., Tronick, S., Aaronson, S.A., Pulciani, S. and Barbacid, M. (1982) T24 human bladder carcinoma oncogene is an activated form of the normal human homologue of BALB and Harvey-MSV transforming genes. Nature 888, 294-298. Schechter, A.L., Stern, D.F., Vaidyanathan, L., Decker, S.J., Drebin, J.A., Greene, M.I. and Weinberg, R.A. (1984) The gag oncogene: an erbB- related gene encoding a 185,000-Mr tumor antigen. Nature 818, 513-516. Schwab, M., Alitalo, K., Klemonauer, K.H., Varmus, H.E., BishOp, J.M., Gilbert, F., Brodeur, G., Goldstein, M. and Trent, J. (1983) Amplified DNA with a limited homology to the myc cellular oncogene si shared by human neuroblastoma cell lines and a neuroblastoma tumor. Nature 888, 245-248. Schwab, M., Varmus, H.E. and Bishop, J.M. (1985) Human N-mys gene contributes to neoplastic transformation of mammalian cells in culture. Nature 818, 160-162. Seeburg, P.H., Colby, W.W., Capon, D.J., Goeddel, D.V. and Levinson, A.D. Comparative biochemical properties of normal and activated human gas p21 protein. Nature 818, 71-75. Sefton, B.H., Trowbridge, I.S., Cooper, J.A. and Scolnick, E.H. The transforming proteins of Rous sarcoma virus, Harvey sarcoma virus and Abelson virus contain tightly bound lipid. Cell 81, 465-474. Shih, C., Shilo, B.Z., Goldfarb, M.P., Dannenberg, A. and Weinberg, R.A. (1979) Passage of phenotypes of chemically transformed cells via transfection of DNA and chromatin. Proc. Natl. Acad. Sci. USA 18, 5714- 5718. 43 Shih, T.Y., Weeks, M., Gross, P., Dhar, R., Oroszlan, S. and Scolnick, E.M. (1982) Identification of the precursor in the biosynthesis of the p21 transforming protein. J. Virol. 18, 253-261. Shimizu, K., Birnbaum, D., Ruley, M.A., Fasano, 0., Suard, Y, Edmund, L., Taparowski, E. Goldfarb, M. and Wigler, M. (1983) Structure of the Ki- gas gene of the human lung carcinoma cell line Calu-l. Nature 888, 497-500. Shimizu, K, Goldfarb, M., Suard, Y., Perucho, M., Li, Y., Kamata, T., Feramisco, J., Stavenzer, E., Fogh, J. and Wigler, M. (1983) Three human transforming genes are related to the viral gas genes. Proc. Natl. Acad. Sci. 88, 2112-2116. Shimizu, K., Nakatsu, Y., Sekiguchi, M., Hokamira, K., Tanaka, K., Terada, M. and Sugimura, T. (1985) Molecular cloning of an activated human oncogene, homologous to v-gas, from a primary stomach cancer. Proc. Natl. Acad. Sci. USA 88, 5641-5645. Sigal, I.S., Gibbs, J.B., D’Alonzo, J., Temeles, G.L., Wolanski, 8.5. (1986) Mutant gas-encoded proteins with altered nucleotide binding exert dominant biological effects. Proc. Natl. Acad. Sci. USA 88, 952-956. Sinn, E., Muller, W., Pattengale, P., Tepler, 1., Wallace, R. and Leder, P. (1987) Coexpression of MMTV/v-Ha-gas and MMTV/c-mgs genes in transgenic mice: Synergistic action of oncogenes in vivo. Cell 18, 465- 475. Spandidos, D.A. and Siminovitch, L. (1978) The reJationship between transformation and somatic mutation. Cell 18, 651-656. Spandidos, D.A. and Wilkie, N.M. (1984) Malignant transformation of early passage rodent cells by a single mutated human oncogene. Nature 818, 469-475. Spandidos, D.A. (1985) Mechanism of carcinogenesis: The role of oncogenes, transcriptional enhancers and growth factors. Anticancer Res. 8, 485-498. Spandidos, D.A., Freshney, M and wilkie, N.M. (1985) Heterogeneity of cell lines derived after transformation of early passage rodent cells by the Ha-gas 1 human oncogene. Anticancer Res. 8, 387-392. Spector, D.H., Varmus, H.E. and Bishop, J.M. (1978) Nucleotide sequences related to the transforming gene of avian sarcoma virus are present in DNA of uninfected vertebrates. Proc. Natl. Acad. Sci. USA 18, 4102-4107. Stehelin, R.V., Guntaka, H.E., Varmus, H.E. and Bishop, J.M. (1976) Purification of DNA complementary to nucleotide sequences required for neoplastic transformation of fibroblasts by avian sarcoma viruses. J. Mol. Biol. 181, 349-354. 44 Steward, T.A., Pattengale, P.K. and Leder, P. (1984) Spontaneous mammary adenocarcinomas in transgenic mice that carry and express MTV/my; fusion genes. Cell 88, 627-637. Streb, H, Irvine, R.F., Berridge, M.J. and Schulz, I. (1983) Release of Ca“ from a nonmitochondrial intracellular store in pancreatic acinar cells by inositol-1,4,5-triphosphate. Nature 888, 67-68. Sukumar, S., Notario, V., Martin-Zanca, D. and Barbacid, M. (1983) Induction of mamary carcinomas in rats by nitroso-methylurea involves malignant activation of H-ras--1 locus by single point mutations. Nature 888, 658-661. Sweet, R.W., Yokoyama, S., Kamata, T., Feramisco, J.R., Rosenberg, M. and Gross, M. (1984) The product of gas is a GTPase and the t24 oncogenic mutant is deficient in this activity. Nature 811, 273-275. Tabin, C.J., Bradley, S.N., Bargemann, C.I., Weinberg, R.A., Papageorge, A.G., Scolnick, E.M., Dhar, R., Lowy, D.R. and Chang, E.H. (1982) Mechanism of activation of a human oncogene. Nature 888, 143-149. Taparowski, E., Suard, Y., Fasano, 0., Shimizu, K., Goldfarb, M. and Wigler, M. (1982) Activation of the T24 bladder carcinoma transforming gene is linked to a single amino acid change. Nature 888, 762-765. Temin, H.M. (1976) The DNA provirus hypothesis: The establishment and implications of RNA-directed DNA synthesis. Science 188, 1075-1080. Temin, H.M. and Mitzutani, S. (1970) RNA-dependent DNA polymerase in the virions of Rous sarcoma virus. Nature 888, 1211-1215. Temin, H.M. (1964) Natl. Caner Inst. Monograph. 11, 557- Thomassen, D.G., Gilmer, T.M., Annab, LA. and Barrett, J.C. (1985) Evidence for multiple steps in the neoplastic transformation of normal and preneoplastic Syrian hamster embryo cells following transfection with Harvey murine sarcoma virus oncogene v-Ha-ras. Cancer Research 88, 726- 732. Thompson, C.B., Humphries, E.H., Carlson, L.M., Chen, C.M. and Neiman, P.E. (1987) The effect of alterations inmgs gene expression on B cell development in the bursa of Fabricius. Cell 81, 371-381. Todaro, G.J. and Huebner, R.J. (1972) The viral oncogene hypothesis: new evidence. Proc. Natl. Acad. Sci. USA 88, 1009-1015. Trahey, M., Milley, RJ., Cole, G., Innins, M, Paterson, H., Marshall, C.J., Hall, A. and McCormick, F. (1987) Biochemical and biological properties of the human N-gas p21 protein. Mol. Cell. Biol. 1, 541-544. Trahey, M. and McCormick, F. (1987) A cytoplasmic protein stimulates normal N-gas p21 GTPase, but does not affect oncogenic mutants. Science 888, 543-545. 45 Treisman, R. (1985) Transient accumulation of c-fos RNA following serum stimulation requires a conserved 5’ element and c-fos 3’ sequences. Cell 18, 889-902. Ullrich, A., Coussens, L., Hayflick, J., Dull, T.J., Gray, A., Tam, A.W., Lee, J., Yarden, Y, Libermann, T.A., Schlessinger, J., Downward, J., Mayes, E.L., Whittle, N., Waterfield, M.D. and Seeburg, P.H. (1984) Human epidermal growth factor receptor cDNA sequence and aberrant expression of the amplified gene in A431 epidermoid carcinoma cells. Nature 888, 418-424. Varmus, H.E. (1984) The molecular genetics of cellular oncogenes. Ann. Rev. Genet. 18, 553-612. Varmus, H.E. (1983) in Mobile Genetic Elements (Shapiro, J.A., ed.), pp. 411-503, Academic Press. Walter, M., Clark, S.G. and Levinson, A.D. (1986) The oncogenic activation of human p21 gas by a novel mechanism. Science 888, 649-652. Waterhouse, J. Muir, C., Correa, P. and Powell, J. (eds.)' (1976) "Cancer incidence in five continents." Lyon, France: IARC Publications No. 15. Willingham, M.C., Pastan, 1., Shih, T.Y. and Scolnick, E.M. (1980) Localization of the sgs gene product of the Harvey strain of MSV to the plasma membrane of transformed cells by electron microsopy. Cell 18, 1005-1014. Willumsen, B.M., Christensen, A., Hubbert, N.L., Papageorge, A.G. and Lowy, D. (1984) The p21 gas C-terminus is required for transformation and membrane association. Nature 818, 583-586. Willumsen, B.M., Papageorge, A.G., Kung, H.J., Bekesi, E., Robbins, T. (1986) Mutational analysis of a gas catalytic domain. Mol. Cell. Biol. 8, 2646-2654. Winter, E., Tamamoto,F., Almoguera, C. and Perucho, M. (1985) A method to detect and characterize point mutations in transcribed genes: Amplification and overexpression of the mutant c-Ki-gas allel in human tumor cells. Proc. Natl. Acad. Sci. 88, 7575-7579. Yuasa, Y, Srivastava, S.K., Dunn, C.Y., Rhim, J.S., Reddy, E.P. and Aaronson, S.A. (1983) Acquisition of transforming properties by alternative point.mutations within c-bas/has human proto-oncogene. Nature 888, 775-779. Zarbl, H., Sukumar, S., Arthur, A.V., Martin-Zanca, D. and Barbacid, M. Direct mutagenesis of Ha-gas 1 oncogenes by N-nitroso-N-methylurea during initiation of mammary carcinogenesis in rats. Nature 818, 382-385. CHAPTER II Morphological Transformation. Focus Formation and Anchorage Independence Induced in Diploid Human Fibroblasts by Expression of a Transfected H-ras OncogeneI Peter J. Hurlin, Dennis 6. Fry, Veronica M. Maher and J. Justin McCormick Carcinogenesis Laboratory, Fee Hall Department of Microbiology and Department of Biochemistry Michigan State University, East Lansing, MI 48824-1316 46 47 Footnotes 1This research was supported in part by the Department of Energy Grant DE-FGOZ-87-ER60524 and by the Department of Health and Human Services Grant CA21289 from the National Cancer Institute. DGF is a recipient of a Leukemia Society of America Special Fellowship Award. 2The abbreviations used are: DMSO, dimethylsulfoxide; EMEM, Eagle's minimal essential medium; FBS, fetal bovine serum; Staph A buffer, 0.2 M phosphate buffer, pH 7.4, l% triton X-lOO, 0.l% SDS, 0.5% sodium deoxycholate, 0.1% NaN3. 0.1 M NaCl. 3Peter J. Hurlin, Dennis G. Fry, Veronica M. Maher and J. Justin McCormick, unpublished data. 48 In an attempt to determine how normal human fibroblasts respond to high expression of the T24 H-ras oncogene, we transfected such cells with the plasmid vector pH06Tl (Spandidos and Wilkie, Nature 3l0:469-475, 1984), containing the T24 H-ras oncogene with 5' and 3' enhancer sequences, and the aminoglycoside phosphotransferase gene which confers resistance to the drug G418. Approximately 1.5% of the G418 resistant colonies obtained after transfection and selection consisted of cells exhibiting obvious morphological transformation, i.e., they were highly refractile, and more rounded than normal fibroblasts. DNA hybridization analysis showed that the morphologically transformed cells contained the transfected T24 H-ras oncogene. and radioimmunoprecipitation analysis showed that they were expressing the T24 H-ras protein product, p21. Morphologically transformed cells formed colonies in soft agar at a frequency at least 60 times higher than that of cells that had been transfected with the control plasmid containing the normal cellular H-ras gene. Cells transfected with plasmid pH06Tl could also be identified by their ability to form distinct foci when grown to confluence in non-selective medium following transfection. This study demonstrates that normal diploid human fibroblasts in culture can be transformed by transfection with a H-ras oncogene. and that such transformation correlates with expression of the mutant p21. 49 IITRINNICTIN A significant pr0portion of human tumors from various sites in the body have been shown to contain activated oncogenes from the ras family (1-3). This finding, which suggests that the ras oncogenes are involved in causing such tumors, has prompted investigations into the role ras oncogenes play in bringing about transformation of cells in culture. Transfection studies using primary or early passage rodent fibroblasts as recipients of ras oncogenes indicate that these genes act in a dominant manner to cause transformation (for review see ref. 4). The transformed phenotypes observed in ras oncogene-transfected rodent fibroblasts include morphological alteration (5.6), focus formation (6), anchorage independence (5,8,9), and, in some instances. tumorigenicity (6-9). Several of these studies indicate that the level of expression of the ras oncogene can be an important factor governing the degree of transformation by ras oncogenes (7-9) and that high expression of the transfected ras oncogene is required for the induction of the tumorigenic phenotype in such fibroblasts (6-9). In contrast to the results achieved using rodent fibroblasts, the results of most DNA transfection studies in which early passage diploid human fibroblasts were used as the recipients suggested that these cells are more resistant to transformation by 31s oncogenes. For example, several groups have reported that following ras oncogene transfection, human fibroblasts failed to exhibit morphological transformation (9,10), focus formation (10), anchorage independence (9), or tumorigenicity (9-ll). Sutherland g_tal. (ll) reported that 50 transfection of human fibroblasts with the T24 H-ras oncogene in a plasmid devoid of transcriptional enhancer sequences caused the target population to exhibit a very high frequency of cells able to form colonies in soft agar. Sager at a_l_. (10,12) reported that following transfection of normal human fibroblasts with the same oncogene, but in a plasmid which contains SV40 transcriptional enhancer sequences. expression of the H-ras p2l could be detected. However. this did not result in morphological transformation or focus formation, or tumorigenicity. These investigators did not assay the transfectants for anchorage independence. An explanation for the negative results with human fibroblasts. suggested by the transfection studies with animal fibroblasts, is that the transfected ras oncogene was not expressed or that the level of expression of the ras oncogene was not high enough to cause measurable effects. In an attempt to achieve high expression of the T24 H-ras oncogene in early passage, diploid human fibroblasts we used a plasmid, pH06Tl, obtained from Spandidos and Hilkie (8), in which the oncogene is inserted between two sets of transcriptional enhancer sequences. We here report that the transfectants exhibited transformation to morphological alteration and focus formation, that progeny of mmrphologically altered colonies exhibited anchorage independence, and that such transformation was correlated with expression of the T24 H-ras oncogene protein product p21 in the human fibroblasts. 5 1 MTERIALS AND m5 Cells and Culture Conditions. Diploid human fibroblasts were initiated from foreskins of clinically normal newborn infants as described previously (13). The human bladder carcinoma cell line T24 was purchased from the American Type Culture Collection (Rockville. MD). EMEMZ, supplemented with 0.2 mM serine, 0.2 mM aspartate, 1.0 mM pyruvate, 10% FBS (Grand Island Biological Co., Grand Island, NY), penicillin (100 units/ml), and streptomycin (lOOug/ml) (complete EMEM) was used for routine culturing of cells. Cells were grown at 37°C, in 5% C02, water saturated incubators. Plasmids. Homer 6 plasmids (8) containing the T24 H-ras oncogene (pH06Tl), the cellular H-ras gene (pH06Nl), or no additional sequences (pH06) were kindly provided by Dr. Neil Wilkie. These plasmids contain the selectable marker gene, aminoglycoside phOSphotransferase, which confers resistance to the drug G418. DIA Transfection. The Polybrene/DMSO method, adapted for transfection of diploid human cells by Morgan _e_t_a_l. (14) was used. Briefly, 1 x 105 cells (passage 7-10) were plated into loo-mm diameter dishes in complete EMEM and incubated at 37°C. After l6-24 h, the medium was removed and replaced with transfection medium, i.e., 2.5 mls of complete EMEM containing 7.5 ug Polybrene per m1 (Sigma Chemical Co., St. Louis, MO) and 1.2-2.0 ug plasmid per ml, and the dishes were returned to 37°C. The transfection medium was distributed over the cells by shaking once every hour. After 6 h the transfection medium was removed and the cells were treated for 4 min at room temperature with 5 mls of complete EMEM containing 30% 52 freshly-distilled DMSO. Following DMSO “shock“ the cells were rinsed 2 times with EMEM, refed with complete EMEM, and incubated at 37°C. Selection of 6418 Resistance. If the transfected population were to be selected for drug resistance, 24 h after DMSO shock the medium was exchanged for complete EMEM containing 200 ug of active G418 (GIBCO) per ml. Colonies were allowed to develop for 14 days. with one refeeding with this selective medium. Assay for Focus Formation. If the transfected cells were to be assayed for the ability to form macroscOpic foci piled up on tap of a monolayer of contact-inhibited cells, they were not selected for G418 resistance, but instead were allowed to grow to confluence in complete EMEM following DMSO "shock'. The cells were fed once per week for three weeks and then stained with methylene blue and the foci counted. Assay for Anchorage Independence. Cells were trypsinzed and resuspended in a modified version of Ham's MCDBllO (15) formulated in this laboratory for use with serum replacements in studies requiring absence of serum (16). For the present assay, 1% FBS was supplied. The cells were counted electronically (Coulter Corp., Hialeah, FL) and plated into top agar at a density of 4,000 cells per ml. The tap agar consisted of the modified MCDBllO medium supplemented with 1% FBS and 0.33% SeaPlaque agar (FMC Corp., Rockland, ME). The bottom agar was prepared with the same medium, but supplemented with 2% agar. One ml of top agar was plated per 35-m diameter dish on tap of 2 ml solidified bottom agar. The following day and weekly thereafter 1 m1 of modified MCDBllO medium with 1% FBS was added to the tap agar to compensate for evaporation. After 3 weeks the 53 number of colonies with a diameter larger than 60 um was determined mi crosc0pi cal 1y. Southern Blot Analysis. Genomic DNA was isolated as previously described (17). DNA was digested with restriction enzymes under conditions reconmIended by the supplier (New England Biolabs, Beverly, MA). Ten ug of digested genomic DNA or 5.6 ug plasmid DNA was electrOphoresed in 0.8% agarose gels and transferred to nylon filters (Gene Screen Plus, New England Nuclear, Boston, MA) using standard techniques (18). Nylon filter prehybridization, hybridization and high strigency washing conditions were carried out as recomended by the supplier. The 2.9 kb _Sas 1 fragment of the T24 H-ras oncogene, nick translated with 3"’P-labeled dNTPs (New England Nuclear, Boston, MA, 600Ci/IImIol) to a specific activity of 1-2 x 108 cpm/ug, was used for hybridization. Filters were exposed to Kodak XAR X-ray film (Kodak, Rochester, NY) with the aid of intensifier screens (Cronex Lightning Plus, Dupont, Inc., Wilmington, DE) for 72 h at -70°C. RNA Analysis. Cyt0p1asmic RNA dot blots were performed as described by White and Bancroft (19). Detection of p21. Protein labeling and iIIInunOprecipitation were carried out essentially as described previously (20,21) with the following modifications. Cells growing exponentially in the modified version of MCDBllO medium supplemented with 10% FBS were plated into Ioo-m diameter dishes at 106 cells/dish. After 24 h the medium was exchanged for 3 ml of this medium lacking methionine and F85, but supplemented with serum replacements (16). After 45 min, 1.0 mCi of 35S-methionine (1148 Ci/Innol, New England Nuclear, Boston, MA) was added and the cells were incubated for 24 h. Cells 54 were lysed with 1.0 m1 Staph A buffer2 containing 2.0 mM phenylmethanosulfonyl fluoride (Sigma Chemical Co., St. Louis, MO) and 135 kallikrein inactivator units/m1 aprotinin (Sigma Chemical Co., St. Louis, MO). Lysed cells were centrifuged at 35,000 rpm in a Beckman 50Ti rotor for 35 min. The supernatant, containing 108 cpm of trichloroacetic acid precipitable counts, was imunOprecipitated for 2 h at 4°C with the monoclonal antibody Y13-259 (22), kindly supplied by Dr. M. Anderson. Protein A-sepharose (Pharmacia Chemical Co., Piscataway, NJ) coated with goat-anti-rat 196 (Cooper Biomedical, Malvern, PA) was added and the mixture was incubated for 30 min at 4°C. The mixture was centrifuged, the supernatant was removed and the sepharose fraction was washed twice with Staph A buffer, once with a solution of 50 mM Tris-Cl (pH 7.5), 1 M MgClz and one more time with Staph A buffer. Electrophoresis sample buffer was added and the samples were heated at 90°C for 5 min. The samples were centrifuged and aliquots of the supernatant, along with pre-stained protein molecular weight markers (Bethesda Research Laboratories, Bethesda, M0), were electrOphoresed (23) through 12.5% SDS-polyacrylamide gels for 4 h at 200 Y. The gels were fixed, treated with autoradiographic enhancer (Enlightning, New England Nuclear, Boston, MA), dried, and exposed to X-ray film at -70°C for 24-72 h. 55 RESULTS Frequency of Transfection to G418 Resistance. Plasmid pHOGTl, which contains the T24 H-ras oncogene, was chosen for transfection of diploid human fibroblasts because of its demonstrated ability to transform early passage rodent fibroblasts into malignant cells (8) and because in our preliminary experiments, other plasmids containing the T24 H-ras oncogene proved to be ineffective in transforming human fibroblasts3. Plasmid pH06N1, which contains the normal, endogenous H-ras gene, and plasmid pHOG» which has no ras gene were used as controls in the transfection experiments. Recombinant plasmids were introduced into early passage (passage 7-10) diploid human fibroblasts using a modified Polybrene/DMSO transfection technique. This transfection technique was chosen because the frequency of G418 resistant clones obtained was shown to be 10 to 40 times that acheived using calcium phosphate (1). To increase the probability that colonies selected for resistance to G418 represented unique transfection events, transfection and selection of cells were performed in the same culture dish. The frequency of G418 resistant colonies obtained after transfection of human fibroblasts with the various Homer 6 plasmids is shown in Table 1. Each plasmid construction yielded approximately the same number of G418 resistant colonies per ug of transfected plasmid, i.e., 200 per 106 target cells. Transformation to Altered Morphology. G418 resistant colonies were microsc0pica11y examined for evidence of altered cellular morphology two wks following transfection and selection in G418 medium. 56 are.» _. nuance—.3 2" «3331322. no 92512. 8:53. mos—.333 2:. «35 3:52.... zcaumw flown. mafia mafia «mummamsn zoxuso.onaomd_k o4 namzmm somamnmsw nodoaamw um: mdnmxma monan cmma a: nodozamm gem nmddm nodozamm um: .om nmddw edemaaa axmsmwmnnaoam amnmnama nswzmqmoamam ems mom flsmamfimnfimn uxmzmwmnema uzom mm muw woo o o uzomzd do _um Hum o o uzomq. dim m.amd mod w um adom nmddm umx .oouaa n‘mamnms mam: tmxm «xmzmqmnfima saw: urn adamaaa cz>. acmamwaazma mines «to swarm om medm Madonnaoa moddozazm «wasmfimnfiaos. nomamwaazma mmfims etc 2mmxm axotas a: noaudmfim mzmz weddozazo axmzmfimnaaos. 57 Figure 1. Comparison of normal and transformed morphologies. (a) normal human fibroblasts; (b) fibroblasts transfected with pH06Tl. Confluent cultures were fixed with methanol and stained with methylene blue. (Magnification, 100x.) 58 Only pHOGTl induced morphological transformation, as indicated in Table 1. A representative colony of morphologically transformed cells is shown in Fig. l. The morphologically transformed cells divided rapidly, were highly refractile and grew in irregular patterns, rather than the highly-oriented pattern of normal human fibroblasts. Transformation to Focus Formation. Cells transformed following transfection with pH06Tl could be also identified by their ability to grow into a 3-dimensional array of cells, a focus, on top of a monolayer of contact-inhibited cells. This type of aberrant growth behavior is a cannon characteristic of tumor cells, including those derived from human fibrosarcomas3. Macroscopic foci, and an individual focus of pH06Tl transfected fibroblasts surrounded by adjacent non-transformed fibroblasts are shown in Fig. 2. The number of foci observed per lOO-Inn diameter dish following three weeks of growth under non-selective conditions was approximately lO-fold greater than the number of morphologically transformed colonies observed after two weeks of G418 selection (see Table 1). No foci were generated following transfection of the control plasmids pH06N1 and pH06 which lack the oncogene. MicroscOpic inspection of the morphology of the cells making up the fbcus revealed that not all cells capable of forming foci exhibited the transformed morphology. Thus, the number of foci is a more sensitive determination of the frequency of transformed cells following transfection with pHOGTl than is the number of morphologically transformed colonies. Induction of Anchorage Independence. Morphologically transformed, G418 resistant colonies obtained following transfection with pH06Tl and G418 resistant colonies obtained following transfection with 59 Figure 2. The focus forming ability of pH06Tl transfected fibroblasts. Following transfection with pHOGTl which contains the T24 H-ras oncogene, or pH06N1 which contains the normal cellular H-ras gene, cells were allowed to grow for 3 weeks under non-selective conditions. Confluent cells were trypsinized and replated at a 1 to 3 dilution and then allowed to grow for an additional 3 weeks before being stained with methylene blue. (a) Foci generated in a dish containing pH06Tl transfectants; (b) confluent monolayer in a dish containing pH06N1 transfectants; (c) 40x magnification of a focus from the dish shown in (a). 6O 61 the control plasmid, pHOGNl, were compared for their ability to form colonies in soft agar. The cell p0pulations assayed were derived from pooled colonies that had been isolated 24 h earlier. No colonies were observed in the control dishes, representing a total of 8,000 cells assayed (frequency, (1.25 x 10'4). In contrast, 122 colonies with a diameter equal to or larger than 60pm were observed with the oncogene transformed cells, out of a total of 16,000 cells assayed (frequency, 76 x 10'4). A concomitant comparison of the ability of the two pepulations to form colonies on plastic showed that the cloning efficiency of the control cells was twice that of the transformed cells (data not shown). Therefore, when measured under the stringent conditions (1% FBS) employed in this assay, cells exhibiting the transformed morphology formed colonies in soft agar at a frequency at least 60 times that of the control cells. Analysis for the Presence of Transfected Sequences. G418-resistant colonies that exhibited the transformed morphologically were isolated, pooled in groups of several clones, and expanded for several population doublings. High molecular weight DNA was extracted and digested with restriction endonucleases Xba I and Mind 111 (double digestion) to liberate specific size fragments (2.5kb and 4.5kb) that are diagnostic for an intact region of plasmid pH06Tl containing the T24 H-ras oncogene, its endogenous promoter and exogenous 5' and 3' enhancer sequences (8). The DNA was analyzed by the method of Southern (18). As shown in Fig. 3, cells exhibiting the transformed morphology contain the 4.5 kb and 2.5 kb fragments. The intensity of these two fragments was approximately half that of the endogenous 62 Figure 3. Southern blot analysis of cellular DNA from morphologically transformed cells obtained following transfection with pH06Tl. Following transfection of pH06Tl, and G418 selection, colonies of morphologically transformed cells were isolated and pooled so that each pool represented 3-4- separate clonal papulations. The clonal populations were expanded and high molecular weight DNA was extracted from them and from the parental strain of normal diploid human fibroblasts used for transfection. lOug of DNA from each sample was digested with Xba I and Hind 111 (double digest), electrOphoresed through a 0.8% agarose gel, and transferred to nylon filters. The filters were hybridized with the 2.9 kb Sac I fragment of the T24 H-ras oncogene radioactively labeled by nick translation. (a) Lane 1, non-transfected normal diploid human fibroblasts; lane 2-5, pooled populations of morphologically transformed cells. Hybridizing fragments of 4.5 and 2.5 kb are diagnostic for the presence of transfected T24 H-ras oncogene sequences. (b) Southern blot analysis of high molecular weight DNA extracted from progeny cells of those analyzed in Fig. 4a, but which had reverted to the normal fibroblast morphology. DNAs were analyzed as above. Lanes 2'-5' are the progeny cells from the respective lanes in Fig. 4a, kb, kilobases. 63 64 ras sequences, indicating that the cells contain only' one capy' of the transfected ras gene. We found that if the morphologically transformed G418-resistant cells were passaged extensively, even in the presence of selective medium, the majority of them eventually reverted to a normal morphology. Of the 38 morphologically transformed colonies analyzed, 36 reverted to a normal morphology within 25 pepulation doublings following transfection. These reverted cells ‘were assayed for 'the ability to form foci, to form colonies in soft agar, and for the presence of the 'T24- H—ras oncogene and its protein product. They did not form foci or colonies in agar. However, Southern blot analysis showed that the revertants still possessed the transfected oncogene (Fig. 4b). Expression of the T24 H-ras Oncogene in Transformed Cells. Substitution of valine for glycine at the 12th amino acid position of the T24 H-ras-encoded p21 alters its electrOphoretic mobility, making it distinguishable from normal, endogenous p21 (24). Seven independent G418 resistant, morphologically transformed clonal p0pulations obtained after pH06Tl transfection and 5 clonal populations that reverted to the normal fibroblast morphology were analyzed for the presence of the mutated, form of p21 by radioimunOprecipitation analysis using monoclonal antibody Y13-259 (22). Fig. 4 shows a comparison of p21 expression in pH06Tl transfectants exhibiting a transformed morphology, and in cells which reverted to a normal morphology. In cells transfected with the control vector (lanes 1 and 7) and in pH06Tl transfectants that reverted (lanes 3-5), the characteristic p21 doublet (20) is expressed (see dashes). The pH06Tl 65 Figure 4w Analysis of ras gene expression. [35S]-methionine labeled cellular extracts were immunoprecipitated with monoclonal antibody Y13-259, the immunOprecipitates electrophoresed on 12.5% polyacrylamide gels, and the gels analyzed by fluorography. Lanes: 1 and 7, cells transfected with the control vector, pH06; 2 and 8, the human bladder carcinomoa cell line T24; 3-5 independent clonal papulations that reverted from a transformed morphology to a normal morphology following transfection with pH06Tl; 6 and 9, pH06Tl transfected, morphologically transformed cells. Lanes 7-9 represent a second analysis of samples 1, 2, and 6, respectively. Dashes: endogenous p21 doublet. Arrows: mutant (T24) H-ras p21. 1 23456 66 789 67 Figure 5. H-ras RNA expression. Cytoplasmic samples were extracted from 8 x 105 cells, denatured, diluted to the equivalent cell number indicated, and spotted onto nitrocellulose filters. (a) Hybridization with the Sac 1 3kb T24 H-ras fragment, nick translated with [32P]-NTPs. Lanes: 1, cells transfected with the control vector, pH06; 2 and 3, cells transfected with pH06Tl and exhibiting a transformed morphology. (b) Hybridization of the same samples as in (a) with a /3-2 microglobulin probe to control for variability in the amount of cyt0plasmic extract loaded. 68 A . 1 2 3 '91 . 2:1105 4 O . 1x105 1“. s 2.5x104 6.25x103 3 1 _2__ 3 g 0 '0' a. 2x105 69 transfectants exhibiting a transformed morphology expressed an additional p21 species that co-migrated with a p21 species expressed in the human bladder carcinoma cell line T24. The mutant species migrated immediately below the slower' migrating, normally present p21 species (see arrow). A similar pattern of expression has been observed in all seven of the morphologically transformed populations analyzed. The additional p21 species was not observed in any' of the reverted papulations, or in control, vector (pH06) transfected p0pulations tested. RNA dot blot analysis of cytoplasmic RNA samples indicated that transfectants maintaining the transformed morphology expressed higher than normal levels of H-ras specific RNA. Fig. 5 shows that samples from such cells (lanes 2 and 3) expressed approximatley four times the amount of H-ras RNA expressed in control cells (vector transfected; lane 1). This higher level of expression of H-ras RNA correlates with expression of the T24-H-ras p21 species in pH06Tl transfected, morphologically transformed cells. TI-origenicity. Two G418-resistant, morphologically transformed clonal populations obtained following transfection with pH06Tl that maintained their altered morphology when expanded into large papulations were assayed for tumorigenicity in athymic mice as described previously (25). Three mice were used per test and 107 cells were injected subcutaneously in each animal. No tumors or indication of growth at the site of injection were observed over a period of greater than one year. When the cells were continuously passaged in culture as part of a lifespan assay, they senesced at a population doubling level comparable to that of normal fibroblast populations. 7O DISCUSSIN Results of the present study show that normal diploid human fibroblasts can be transformed into morphologically altered, focus forming, and anchorage independent cells following transfection of the T24 H-ras oncogene. Our results, showing that the transformed cells exhibited anchorage independence, support the results obtained by Sutherland e_t_ _a_1_ . (11) and extend that study by showing that the transformed cells expressed the T24 H-ras protein. However, unlike Sutherland fl; al. (11) we did not use anchorage independence to identify transformed papulations of cells directly following transfection. Therefore, it is not possible to directly compare our frequency of anchorage independent cells with theirs. If the anchorage independence exhibited by cells transfected with the 124 H-ras oncogene in the enhancer-free plasmid used by Sutherland _e_t_ g. (11) is an indication that the transfectants were expressing the mutant p21 protein, then it is difficult to understand why Sager _e_t_ al, (10,12), did not observe any evidence of transformation. They used the same oncogene, but in a plasmid which contains a set of transcriptional enhancer sequences, and they found expression of mutant p21. An obvious explanation for the difference between the results of Sager _e_t_ _a_l. (10,13) and those we present here is a difference in the level of expression of p21 resulting from the use of different plasmid constructions containing the 124 H-ras oncogene. When we began transfection studies using normal diploid human fibroblasts as recipients, like Sager gal. (10,12) we constructed and used a pSV2-derived plasmid containing the T24 71 H-ras oncogene and found, as they did, no evidence of morphological transformation or focus formation3. What is more, there was no induction of anchorage independence3. The positive results reported in the present study were generated using a plasmid (pHOGTl) that was specifically designed by Spandidos and Wilkie (8) to elicit high expression of the T24 H-ras oncogene in mammalian cells by inserting the oncogene between two sets of transcriptional enhancer sequences. Therefore, we interpret our results to indicate that a sufficiently high level of expression of the T24 H-ras oncogene was achieved to cause the transformation of normal human fibroblasts, and that the pSV2-T24 plasmid of Sager _et _a_l_. (12) was not capable of generating such levels. This interpretation does not explain the negative results of Spandidos (9), who recently reported that transfection of normal human fibroblasts with pH06Tl did not cause anchorage independence or any obvious morphological transformation. However, Spandidos used a different method of transfection (calcium phosphate) then we did (DMSO/Polybrene), and examined a relatively small number of independent G418 resistant colonies (ni200) compared to the total number we examined (>2461). The results of numerous ras oncogene transfection studies carried out with rodent fibroblasts suggest that high expression of the T24 H-ras oncogene can be attained, and that with such expression the tumorigenic phenotype is induced (6-9). Our studies with human fibroblasts indicate that the transfected oncogene is expressed at low levels, compared to what has been observed in rodent fibroblasts, and that such expression is lost in most cells as they are passaged. Only two papulations of pH0611 transfectants out of 38 analyzed could 72 be isolated that maintained a transformed morphology when extensively passaged, and both of these proved to have a finite lifespan and were not tumorigenic. However, there is evidence that ras oncogenes are involved in ne0p1astic transformation of human fibroblasts. For example, transfection of diploid fibroblasts derived from Bloom's syndrome patients with DNA from NIH 3T3 cells that had been transformed by Harvey murine sarcoma virus (Ha-MSV) DNA resulted in cells able to form nodules in athymic mice (26). These nodules subsequently regressed (26). Also, Harvey (27) or Kirsten murine sarcoma virus (28) infection of human fibroblasts that had acquired an indefinite lifespan in culture following SV40 virus transformation or repeated exposure to ganlna radiation resulted in cells able to form tumors. One group (27) indicated that the tumors regressed, the other (28) found they did not. Three groups (29-31) have succeeded in causing ne0plastic transformation of human epithelial cells by ras oncogene transfection. Further studies to determine if diploid human fibroblasts that continue to express the T24 H-ras oncogene are tumorigenic and whether this gene cooperates with other* oncogenes (e.g., myc) in causing malignant transformation of human fibroblasts are in progress. 73 ACKIOHLEDGEMENTS We thank Dr. Neil Wilkie for kindly supplying us with the pH06 series of plasmids and Dr. Marshall Anderson for the Yl3-259 antibody to H-ras p21, and Carol Howland for typing the manuscript. REFERENCES Fugita, J., Yoshida, 0., Tusas, Y., Rhim, J. S., Hatanaka, M., and Aaronson, S. A. Ha-_I_‘_a_s oncogenes are activated by somatic alterations in human urinary tract tumours. Nature, 389: 464-466, 1984. Eva, A., Tronick, S. R., Gol, R. A., Pierce, J. H., and Aaronson, S. A. Transforming genes of human hematOpoietic tumors: Frequent detection of gas-related oncogenes whose activation appears to be independent of tumor phenotype. Proc. Nat. Acad. Sci. USA, g9: 4926-4930, 1983. Santos, E., Martin-Zanca, D., Reddy, E. P., Pierotti, M. A., Della Porta, G., and Barbacid, M. J. Malignant activation of a K-ras oncogene in lung carcinoma but not in normal tissue of the same patient. Science, _2_2_3_: 661-668, 1984. Balmain, A. Transforming ras oncogenes and multistage carcinogenesis. British J. Cancer, 8: 1-7, 1985. Land, H., Parada, L., and Weinberg, R. A. Tumorigenic conversion of primary embryo fibroblasts requires at least two c00perating oncogenes. Nature, 80_4: 596-602, 1983. Pozzatti, R., Muschel, R., William, J., Padmanabhan, R., Howard, B., Liotta, L., and Khoury, G. Science, g8: 223—227, 1986. Land, H., Chen, A. C., Morgenstern, J. P., Parada, L. F., and Weinberg, R. A. Behavior of myc and ras oncogenes in transformation of rat embryo fibroblasts. Mol. Cell. Biol. 8: 1917-1925, 1986. 74 IO. 11. 12. 13. 14. 75 Spandidos, D. A., and Wilkie, N. M. Malignant transformation of early passage rodent cells by a single mutated human oncogene. Nature, 8E: 469-475, 1984. Spandidos, D. A. Mechanism of carcinogenesis: The role of oncogenes, transcriptional enhancers and growth factors. Anti. Can. Res. _5_: 485-598, 1985. Sager, R., Tanaka, K., Lau, C. C., Ebina, Y., and Anisowicz, A. Resistance of human cells to tumorigenesis induced by cloned transforming genes. Proc. Natl. Acad. Sci. USA, 88: 7601-7605, 1983. Sutherland, B., Bennet, P. V., Freeman, A. G., Moore, S. P., and Strickland, P. T. Transformation of human cells by DNA transfection. Proc. Natl. Acad. Sci. USA, 3: 2399-2403, 1985. Sager, R. Resistance of human cells to transformation. 13;: G. F. Y. Woude, A. J. Levine, W. C. Topp, and J. 0. Watson (eds.), Cancer Cells: Oncogenes and Viral Genes, Vol. 2, pp. 487-493. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 1984. McCormick, J. J., and Maher, V. M. Measurement of colony-forming ability and mutagenesis in diploid human cells. 1'1: E. C. Friedberg and P. A. Hanawalt (eds.), DNA Repair: A Laboratory Manual of Research Procedures, pp. 501-522, New York: Marcel Dekker, 1981. Morgan, T. L., Maher, V. M., and McCormick, J. J. A procedure of high efficiency DNA-mediated gene transfer in normal human fibroblasts. In Vitro Cell. and Develop. Biol. 88: 317-319, 1986. 15. 16. 17. 18. 19. 20. 21. 76 Bettger, W. J., Boyce, S. T., Walthall, B. J., and Ham, R. C. Rapid clonal growth and serial passage of human diploid fibroblasts in a lipid-enriched synthetic medium supplemented with epidermal growth factor, insulin, and dexamethasone. Proc. Natl. Acad. Sci. USA, 18, 5588-5592, 1981. Ryan, P. A., McCormick, J. J., and Maher, V. M. Modification of MCDBllO medium to support prolonged growth and consistent high cloning efficiency of human diploid fibroblasts. Exp. Cell Res. .118, 318-328, 1987. Heflich, R. H., Dorney, D. J., Maher, V. M., and McCormick, J. J. Reactive derivatives of benzo[a]pyrene and 7,12-dimethylbenz[a]anthracene cause S] nucleases sensitive sites in DNA and “UV like" repair. Biochem. Bi0phys. Res. Commun. 115 634-641, 1977. Southern, E. M. Detection of specific sequences among DNA fragments separated by gel electrophoresis. J. Mol. Biol. fl: 503-517, 1975. White, B. A., and F. C. Bancroft. Cytoplasmic dot blot hybridization. J. Biol. Chem. 881; 8569-8572, 1982. Der, C. J., and C00per, G. M. Altered gene products are associated with activation of cellular r_ask genes in human lung and colon carcinomas. Cell_88: 201-208, 1983. Reynolds, S. H., Stowers, S. J., Maronpot, R. R., IAnderson, M. W., and Aaronson, S. A. Detection and identification of activated oncogenes in spontaneously occurring benign and malignant hepatocellular tumors of the 8603F1 mouse. Proc. Natl. Acad. Sci. USA, 88; 33-37, 1986. 22. 23. 24. 25. 26. 27. 28. 29. 77 Furth, M. E., Davis, L. J., Fleurdelys, and Scolnick, E. M. Monoclonal antibodies to the p21 products of the transforming gene of Harvey murine sarcoma virus and a cellular g_a_s_ gene family. J. Virol. 13: 294-304, 1982. Laemmli, V. K. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature (London), 2_27: 680-685. 1970. Tabin, C. J., Bradley, S. M., Berman, C. 1., Weinberg, R. A., Papageorge, A. G., Scolnick, E. M., Phar, R., Lowy, D. R., and Chang, E. H. Mechanism of activation of a human oncogene. Nature (London), 80_O_: 143-149, 1982). Fry, 0. G., Milam, L. D., Maher, V. M., and McCormick, J. J. Transformation of diploid human fibroblasts by DNA transfection with the v-s_i_s oncogene. J. Cell. Physiol. 188: 313-321, 1986. Doniger, J., Dipaolo, J. A., and Popesco, N. C. Transformation of Bloom's syndrome fibroblasts by DNA transfection. Science, 828: 1144-1146. 1983. Namba, M., Nishitani, K., Fukushima, F., Kimoto, T., and Nose, K. Multistep process of ne0plastic transformation of normal human fibroblasts by 60Co gaInna rays and Harvey sarcoma viruses. Int. J. Cancer, _31: 419-423, 1986. O'Brien, W., Stenman, G., and Sager, R. Suppression of tumor growth by senescence in virally transformed human fibroblasts. Proc. Natl. Acad. Sci. USA, 88: 8659-8663, 1986. Yoakum, G. H., Lechner, J. F., Gabrielson, E., Korba, B. E., Malan-Shibley, L., Willey, J. C., Valerio, M. G., Shamsuddin, A. K. M., Trump, B. F., and Harris, C. C. Transformation of 30. 31. 78 human broncial epithelial cells transfected by Harvey r_as oncogene. Science, _2_2_Z_: 1174-1179, 1985. Rhim, J. S., Jay, G., Arnstein, P., Price, F. M., Sanford, K. K., and Aaronson, S. A. Neaplastic transformation of human epidermal keratinocytes by A012-Sv-40 and Kirsten sarcoma viruses. Science, _2_2_7_: 1250-1252, 1985. Boukamp, P., Stanbridge, E. J., Cerutti, P. A., and Fusenig, N. E. Malignant transformation of 2 human-skin keratinocyte lines by Harvey-gas oncogene. J. Inves. Der., g: 131, 1986. CHAPTER III Malignant Transformation of a Human Fibroblast Cell Strain, MSU-1.l, By Expression of a Transfected T24 H-gas Oncogene Peter J. Hurlin, Veronica M. Maher and J. Justin McCormick Carcinogenesis Laboratory, Fee Hall Department of Biochemistry and Department of Microbiology Michigan State University, East Lansing MI, 48824-1316 79 80 SUMMARY We showed previously that normal diploid human fibroblasts that express a transfected 124 H-gas oncogene exhibit several characteristics of transformed cells, but are not tumorigenic. To extend these studies of the transforming ability of the 124 H-gas_oncogene in human cells, we have utilized an infinite lifespan, but otherwise phenotypically normal human fibroblast cell strain, MSU-1.1, which was developed in this laboratory following transfection of diploid fibroblasts with a plasmid carrying a v-mys oncogene. Transfection of MSU-1.1 cells with the 124 H-gas oncogene flanked by 2 transcriptional enhancer elements (pH0611) caused foci of morphologically transformed cells to develop. No such transformation occurred if the plasmid containing the 124 H-gas oncogene had only one enhancer or none at all, or if the normal H-ras gene was transfected in the pH06 vector. Cells derived from such foci expressed high levels of 124 H-ras p21, formed colonies in soft agar at high frequency, proliferated rapidly in serum-free medium1containing 0.1 mM calcium, which does not support growth of the parental cell line, and formed progressively growing, invasive fibrosarcomas in athymic mice. No karyotypic changes accompanied the transformation of MSU-l.l fibroblasts, suggesting that chromosome instability' did not play a role in the conversion of the cells to malignancy. Transfection of two other infinite lifespan fibroblast cell lines with the T24 H-gas oncogene in pH06Tl also resulted in malignant transformation, suggesting that the infinite lifespan phenotype of MSU-l.l cells, and not necessarily expression of the v-myc oncogene, was the factor that complemented 124 H-ras expression to cause malignant transformation. 81 INTRODUCTION Oncogenes of the gas family have been detected in numerous and diverse human malignancies. A growing body of evidence from both in vivo and in vitro studies suggests that the p21 products of gas oncogenes are involved in the genesis of such malignancies. For example, transgenic mice harboring the 124 H-gas (1), or v-H-gas oncogenes (2) linked to a murine mammary protein gene promoter developed mammary tumors following a long latency. 'These tumors were clonal in origin, suggesting that additional changes at the cellular level were required to cause tumor formation. Clues to the nature of the steps or events that may complement mutational activation of gas genes in tumorigenesis have come primarily from transfection experiments utilizing rodent fibroblasts in culture (for review see ref. 3). When primary or early passage rodent fibroblasts are transfected with a combination of a ras oncogene and any one of a number of genes thought to be involved in causing cellular imortalization, malignant transformation results (4-6). Transfection of gas oncogenes into infinite lifespan rodent fibroblast cell lines usually results in malignant transformation (4,7,8), but not in all cases (9). However, several studies have reported that high levels of gas oncogene expression can convert finite lifespan, early passage rodent fibroblasts into infinite lifespan malignant cells without the need for transfection of another oncogene (IO-12). To test models proposed for r_as transformation generated using rodent fibroblasts for their relevance to human cell transformation, we have been using human fibroblasts as recipients of transfected ,gas oncogenes (13,14). Several groups reported that early passage diploid human 82 fibroblasts are resistant to transformation by transfected gas oncogenes (15,16), or microinjection of gas oncogene encoded p21 (17,18). But, Hurlin et. al. (13) showed that if such cells are transfected with the pH0611 plasmid which contains the 124 H-gas oncogene flanked by Moloney murine sarcoma virus LTR and SV40 virus transcriptional enhancer sequences, they become morphologically transformed, develop distinct foci, and form colonies in soft agar. In a related study, Tubo and Rheinwald (19) showed that diploid human fibroblasts transfected with the 124 H-gas oncogene in a plasmid which contains an SV40 virus enhancer become morphologically transformed and exhibit reduced growth factor requirements. However, the transformed cells in these two studies did not acquire an infinite lifespan and were not tumorigenic (14,19). For the most part, oncogenes have been detected and defined by their transforming effects on rodent fibroblasts which have an infinite lifespan in culture, but are otherwise normal. No comparable human cell line has been available for use in testing the applicability of the results to human fibroblasts. This is because human fibroblasts in culture are extremely difficult to transform into cells with an infinite lifespan (20,21), even by SV40-virus infection (22) and the few infinite lifespan, non-tumorigenic fibroblast cell lines that have been generated (22-24) consist of cells that are morphologically transformed and exhibit other transformation characteristics. Recently, a human fibroblast cell strain with an infinite lifespan in culture, designated MSU-l.1, was generated in this laboratory in experiments examining the properties induced in finite lifespan, diploid human fibroblasts by transfection of a plasmid carrying a v-mgs oncogene. MSU-l.1 cells have a normal morphology and a stable, near diploid karyotype. They do not form foci, or produce colonies 83 in soft agar, and are non-tumorigenic. Details on the generation of this cell strain, and its properties will be described eleswhere. In the present study we show that transfection of MSU-1.l cells with the 124 H- ras oncogene in pH06Tl caused morphological transformation and focus formation and that cells derived from such foci exhibited reduced growth factor requirements, formed colonies in soft agar at high frequency and gave rise to rapidly growing, malignant fibrosarcomas in athymic mice. 84 METHODS AND MATERIALS Cells and Culture Conditions. Cell line LG1, from which the MSU-1.l cell strain arose was initiated from foreskin material of a clinically normal newborn male as described (25). The infinite lifespan cell strain MSU-1.1, described above, is resistant to G418 (Grand Island Biological Co., Grand Island, NY) because the plasmid used to transfect the original LGl population carried the aminoglycoside phosphotransferase gene. The infinite lifespan human fibroblast cell lines KMST-6 and SUSM-I, which were generated following repeated carcinogen treatment (23,24), were obtained from M. Namba. The SV40 virus transformed, infinite lifespan human fibroblast cell line GM0637B was purchased from the NIGMS Human Genetic Mutant Cell Repository, Camden, NJ. The human bladder carcinoma cell line 124 was purchased from the American Type Culture Collection (Rockville, MD), and the chicken hepatoma, MC29 virus-producer cell line LSCC-DU72 was provided by K. Nazerian. The cells were routinely cultured in Eagle’s minimum essential medium, supplemented with 0.2 mM serine, 0.2 mM aspartate, 1.0 mM pyruvate, 10% fetal bovine serum (FBS) (Gibco), 100 units/ml penicillin, and 100 ug/ml streptomycin at 37°C, in 5% C02, in water saturated incubators. For studies of growth factor independence, a serum-free medium, McM, a version of MCDBllO (26) formulated for use with serum-replacements (27), was employed. Plasmids and DNA Transfection. Homer 6 plasmids (10) containing the 124 H-gas oncogene (pH0611), the normal human H-yas gene (pH06N1) or no additional sequences (pH06), were kindly provided by N. Wilkie. Plasmid p124 was purchased from the American Type Culture Collection, and plasmid 85 pSV2-124 was constructed in this laboratory by inserting the T24 H-gas oncogene into the BamHI site of the plasmid pSV2neo (28).'arget cells (2 x 105) were transfected with plasmid DNA at 1.3 ug per ml in transfection medium, using the Polybrene/dimethylsulfoxide method (29) as previously described (13). Assay for Growth Factor Independence. Cells (5 x 10‘) were plated in McM medium supplemented with 1% FBS. After 24 h, the number of cells was determined by electronic counting and the medium was exchanged for McM medium containing 0.1 mM calcium and supplemented with the serum replacements of Ryan et al. (27), but lacking epidermal growth factor. The medium was renewed 3 days later, and the number of cells was determined after another 3 days. Assay for Anchorage Independence. Cells were assayed for the ability to form colonies in soft agar as described (13) with the following modifications. Cells (5000 per dish) were plated in top agar consisting of McM medium supplemented with 2% FBS and 0.33% SeaPlaque agar (FMC Corp., Rockland ME.) McM medium containing 2% FBS was provided to cultures weekly. Colonies having a diameter larger than 80 um*were counted and sized electronically (Olympus Cue-2 Image Analyser, Olympus Corp. Lk. Success, NY) after 3 wk. Tumorigenicity Assays. The ability of cells to form tumors was assayed by injecting 1 x 107 exponentially growing cells subcutaneously into 4 wk to 8 wk old BALB/c athymic mice which had been irradiated with 300 rads from a °°Co source 24 h previously. To test for the ability to metastasize, 86 5 x 10‘ cells were injected into the tail vein of athymic mice, and after 8 mo the mice were sacrificed and complete necropsies were performed. DNA Analysis. DNA was isolated as previously described (30) and analysed by the method of Southern (31) for transfected sequences. Analysis of gas p21 and v-mys p110. Cells to be analysed for expression of r_as p21 were labeled with [3SS]-Met (New England Nuclear, Boston, MA) at 250uCi per ml for 18 h in McM medium lacking methionine, but supplemented with 1% FBS. Cell lysis, immunoprecipitation with the gas p21 specific antibody Y13-259 (32), NaDodSOg-PAGE and fluorography were performed as previously described (13). For analysis of v-mys p110, cells were labeled as above, but for only 1.5 h. Cells were washed once in ice- cold phosphate-buffered saline, lysed in lysis buffer (10 mM Tris hydrochloride (pH 8.2), 150 mM NaCl, 1% sodium deoxycholate, 1% TRITON X- 100, 0.1% NaDodSO‘, 5 mM EDTA, and 2 mM phenylmethanosulfonyl fluoride), and clarified by centrifugation at 10,000 x g for 20 min. Supernatants containing 7.5 x 107 cpm were immunoprecipitated with a sheep antibody (Cambridge Research Biochemicals, Valley Stream, NY, antibody OA-11-801) that reacts broadly with the proteins of the guys family. Protein A- Sepharose (Pharmacia Chemical Co., Piscataway, NJ) coated with goat anti- sheep 196 (Cooper Biomedical, Malvern PA) was added, and the mixture was shaken for 30 min at 49C. The Sepharose fraction was washed twice with lysis buffer and once with 10 mM Tris hydrochloride (pH 7.5), 1 M MgClZ, and one more time with the lysis buffer. NaDodSO‘-PAGE sample buffer containing 8 M urea was added, and the samples were heated for 5 min at 95°C. The samples were centrifuged, and aliquots of the supernatant, along 87 with prestained molecular weight markers (Bethesda Research Laboratories, Bethesda, M0) were electrophoresed (33) in 10% NaDodSO4-polyacrylamide gels. The gels were fixed, treated with autoradiographic enhancer (ENHANCE, NEN, Boston, MA), dried, and exposed to X-ray film at -80°C. Karyotype1Analysis. G-banded chromosomes were prepared as described (34). For each cell line, at least 25 G-banded karyotypes were prepared to examine for chromosomal rearrangements, and 100 conventially stained metaphases were counted to determine the modal chromosome number. 88 RESULTS Transformation to Focus-Formation by the 124 W-gas Oncogene. Following transfection of MSU-1.1 fibroblasts with the various plasmids, the cells were cultured for 3 wk in Eagle’s medium containing 10% FBS, and screened for colonies of cells that overgrew the monolayer (focus formation). With cells that had been transfected with pH06Tl, distinct foci of morphologically altered cells were found at a frequency of approximately 10 per 10° cells transfected. No foci, or indications of transformation were observed following transfection with control plasmids containing either the normal human H-gas gene (pH06N1), or lacking a gas gene (pH06). Transfection of 5 x 10° MSU-1.1 cells with plasmids containing the 124 H- gas oncogene, but no enhancer sequences (p124), or with only one SV40 transcriptional enhancer (pSV2-124) did not induce focus-formation. Six populations of morphologically transformed fibroblasts (designated MSU-1.1-124 strains 1-6) were isolated from foci in separate dishes, and the populations were expanded for subsequent characterization. An example of the morphology of MSU-1.1-124 cells, which are smaller, rounder and more refractile than their parental cells, is shown in Fig. 1. These cells have maintained their transformed morphology for more than 40 population doublings thus far. This stability contrasts with what Hurlin et a1. (13) found with finite lifespan cells transformed following transfection of the 124 H-gas oncogene in pH06Tl. Growth of MSU-1.1-T24 Fibroblasts in Serum-Free Medium. Schilz et. a1. (35) showed that human fibrosarcoma-derived cell lines, but not normal diploid human fibroblasts, were capable of proliferating in McM medium 89 .32 838:282 .83 83395: m 528. 42-2-3: ea 28 8:329... 3.3: 25 82:63: 5.5:. 2: do 3268:: ._ 85m: .x W.,! . I ‘6 t r .. 1...} flir. \WIC «fin . o. .e ( xii ( #9.. ’0. Q .7. twee —\.\ P \x. . .10.. .1 .\ MIL,- . “.11 I 1 \\dem\0w. .mMIIMIIl .V\_ .\ x \x) . «with. 90 containing 0.1 mM calcium, instead of the usual 1 mM, and supplemented with the serum replacements of Ryan et. al. (26), but lacking epidermal growth factor. The six MSU-1.1-124 cell strains, the parental MSU-1.1 cells and the LGl cell line from which the latter arose were compared for their ability to proliferate in such medium. MSU-1.1 cell strains proliferated to numbers 6 to 7 fold higher than the parental MSU-1.l cells did, and as expected, LG] cells could not proliferate under these conditions (Table 1). The number of cells attained by the MSU-l.1-124 cell strains indicated that, under these conditions, they double at a rate approximately equal to that achieved in medium containing 10% FBS, i.e., once every 24 h. Anchorage Independence of MSU-1.1 124 Cell Strains. The six MSU-1.l-124 cell strains were tested for their ability to form colonies in 0.33% agar medium. LG] cells, and the parental MSU-1.1 cell strain formed colonies with an efficiency of less than 0.007%, i.e., no colonies were detected out of a total of 15000 cells plated. The six MSU-1.1-124 cell strains formed colonies in agar at efficiencies ranging from 3.5% to 23% (Table 1). There was variation in the size range of colonies detected. For example, although MSU-1.1-124 strain 6 exhibited a cloning efficiency of only 3.5%, it gave rise to the highest proportion of very large colonies (i.e., having diameters of >300um). Tumorigenicity of MSU-1.1-124 Cell Strains. Four of the six MSU-I.1-124 cell strains tested for the ability to form tumors in athymic mice formed progressively growing tumors (Table 1). Four to 5 wk were required for the tumors to reach 1 cm in diameter. They were diagnosed as 91 Figure 2. Histology showing a poorly-differentiated human fibrosarcoma formed in an athymic mouse by MSU-1.1-124 strain 3 cells. A 2.5 cm diameter tumor was removed 8 wk following subcutaneous injection of 107 cells, and was sectioned, and stained with hematoxylin-eosin. Mouse skeletal muscle is indicated by the arrows. Magnification; (A) 40x, (8) 200x. 92 93 Table 1. Growth properties and tumorigenicity of MSU-1.1-124 cell strains Cell no. Cloning attained in Days for efficiency serum-free tumor to Cell in agar 0.1 mM Ca++ Tumor reach strain (%) medium (x10'5)a incidenceb 1 cm diam 1 15 16 2/2 36 2 23 NTc 6/6 30 3 10 NT 6/6 30 4 18 17 0/5 -- 5 8.5 19 0/3 -- 6 3.1 17 4/4 34 MSU-1.1 (0.01 2.7 0/9 -- LG1 cells <0.01 0.4d 0/9 -- aDuplicate cultures were counted 7 d after plating 5 x 104 cells/dish. bAnimals with tumors per animals injected subcutaneous. Animals were examined for tumor formation for at least 8 mo following injection. cNT, not tested. dThis is the same no. of cells as was determined 1 d after plating. 94 poorly-differentiated, invasive human fibrosarcomas with a high mitotic index and a moderate degree of anaplasia. A low proportion of giant multinucleated cells were observed in several of’ the tumors. A histological section of a representative tumor is shown in Fig. 2. Two criteria indicated that the tumors were formed by the injected MSU-1.1- 124 fibroblasts. First, the tumor-derived cells that were returned to culture were capable of proliferating in medium containing G418, and second, they contained the marker chromosomes characteristic of MSU-1.1 cells (see below). One of the tumorigenic MSU-1.l-TZ4 strains was tested for metastatic properties by being injected into the tail vein of mice. No tumors were found from necropsies performed on 5 such mice 8 months later. Presence of 124 H- as and v-mys_0ncogenes. Following digestion of genomic DNA with 18a 1 and glad III, fragments of 4.5 and 2.5 kb that hybridize with the 8as I, 2.9 kb fragment of the 124 H-gas oncogene are diagnostic for plasmid pH0611 sequences which include the 124 H-gas oncogene and flanking 5’ and 3’ transcriptional enhancer sequences (10). As shown in Fig. 3, genomic DNA from each of the MSU-1.1-T24 cell strains contained H-gas_hybridizing fragments close to these predicted sizes. Although the data are not shown, v-mys hybidizing DNA fragments were detected in the six MSU-1.1-124 cell strains and they'were of identical size and intensity as those detected in the parental MSU-1.1 cell strain, indicating that no rearrangements or amplification of this oncogene had occurred subsequent to introduction of plasmid pH06Tl. Expression of 124-H-gas and v-mys Oncogenes. Radioimunoprecipitation analysis of the MSU-1.1-124 cell strains (Fig. 4) using the r_as p21 95 1.0.1 . MSU—l.’l MSU-1.1-T24 Cell Strains 2 Figure 3. Southern blot analysis of plasmid pH0611-derived 124 H-gas oncogene sequences in MSU-1.1-124 cell strains. LGl is the diploid human fibroblast cell line from which the MSU-1.1 cell strain was derived. DNA was digested with 818d III and 18a I and hybridized with the 2.9 kb 8as I fragment of the T24 H-gas oncogene. The locations of the expected 4.5 kb and 2.5 kb plasmid pH0611 specific fragments are indicated. 96 Figure 4. Expression of 124 H-gas p21 in MSU-1.1-T24 cell strains. The gas encoded p21 proteins were immunoprecipitated with antibody Y13-259 (25) from f”$]methionine-labeled cell lysates and analysed by electrophoresis in 12.5% SDS-NaDod504 polyacrylamide gels and flourography. The location of the endogenous gas encoded p21 protein doublet is indicated by the bars, and the 124 H-gas encoded p21 location is indicated by the arrow. 124 is the human bladder carcinoma cell line. 3a is a clonal derivitive of MSU-1.1-124 strain 3. The location of molecular weight (kilodaltons) markers on the gel is indicated. 97 'T‘ a a :1. MSU- 2.1 1—1‘24 cell strains .4 E E- 1 4 3 3a 42 - - ‘1 3 '5 H . a. 1 " ‘ J‘ . 25- 98 Ms_u-1.1-Tz4 Cell Strains 1 2 3 4 5 6 LSCC-DU72 MSU-l.l H ‘5 _l 206 - 100 - 42 - Figure 5. Expression of v-myg encoded p110 in MSU-1.1-T24 cell strains. v-myg_p110 was immunoprecipitated fronl[35$]methionine-labeled cell lysates with an antibody that reacts with proteins of the my; family and analysed by electrophoresis in 10% polyacrylamide gels and fluorography. The location of the v-myg encoded p110 is indicated by the arrow. LSCC-DU72 is a chicken hepatoma, M629 virus-producer cell line. The location of molecular weight (kilodalton) markers on the gel is indicated. 99 specific monoclonal antibody Yl3-259 (28), indicated that each of them overexpressed a 21 kd protein (see arrow) that migrated to the same position as the overexpressed T24 H-[as oncogene encoded p21 of the T24 bladder carcinoma cell line. To determine the relative levels of_ra§ p21 expression, the radiolabeled p21 bands were cut out of duplicate gels and quantitated by scintillation counting. The level of total [as p21 expression in the MSU-l.l-T24 cell strains was 3 to 5 times that expressed in the parental MSU-l.l cell strain, or in the LG] cell line from which the latter was derived, and was only slightly elevated over that observed in T24 bladder carcinoma-derived cells. Radioimmunoprecipitaion analysis of [as p21 was also carried out on MSU-l.l—T24 fibroblast cell strains derived from tumors generated in athymic mice. T24 H-Las p21 levels in these cells were comparable to levels determined in their parental focus-derived cells prior to injection into athymic mice, indicating that no in vivo selection of cells expressing higher or lower levels of the transfected oncogene had occurred (data not shown). Radioimmunoprecipitation analysis was also used to compare the level of v-myg p110 expression in the MSU-l.l-TZ4 cell strains, with that in the parental, MSU-l.l cell strain, using as a positive control a chicken hepatoma cell line LSCC-DU72 that expresses a v-myg p110. Such analysis indicated that in the MSU-l.l-T24 cell strains, v-myg p110 was expressed at levels comparable to that found in the parental MSU-l.l cells (Fig. 5, see arrow). As expected, no v-myg p110 was detected in the LEI cells. Karyotype Analysis of USU-1.1-T24 Cell Strains. Analyses of G-banded, and conventionally stained chromosomes from MSU-1.l fibroblasts and 100 MSU-l.l-T24 fibroblasts were carried out to determine karyotype and chromosome stability respectively. A detailed description of the banding patterns for MSU—l.l fibroblast chromosomes will be reported elsewhere. The MSU-1.l-T24 cell strains analysed (strains 2 and 3), as well as cells taken from the tumors derived from them, showed the same karyotype and modal chromosome number, i.e., 45, as the parental MSU-l.l cell strain, as well as the two marker chromosomes characteristic of this cell line. Tumorigenic Transformation of Infinite Lifespan Human Fibroblasts Derived from Other Cell Lines. To determine whether expression of T24 H-ras in infinite lifespan, non-tumorigenic human fibroblast cell lines other than MSU-l.l also caused tumorigenic transformation, we transfected pH06Tl, or the control plasmid, pH06N1, into the SUSM-l (23) and KMST-G (24) cell lines, as well as into a cell line (GM0637B) generated following SV40 virus infection. The transfectants were selected for 6418 resistance, rather than focus formation, because unlike MSU-l.l cells, these target cell lines already exhibit a transformed morphology and aberrant growth behavior. G418-resistant colonies were pooled, and their progeny cells tested for tumorigenicity in athymic mice. Within 4.5 mo, four out of five athymic mice injected with pH06Tl-transfected KMST-6 cells (KMST-G-T24) and three out of six injected of pH06Tl-transfected GM0637B cells (GM0637B- T24) developed progressively growing malignant fibrosarcomas within that reached a 1 cm diameter. No tumors were formed in 5 mice injected with pH06Tl—transfected, SUSM-l cells (SUSM-l-T24) generated in two separate experiments, or' with any of the pHOGNl transfected, G418-resistant controls. Because of the long time required for tumor development by these 101 cell strains relative to MSU-1.1—T24 tumorigenic cell strains, cells were cultured from tumors formed by these cell strains and reinjected in athymic mice to determine if an in vivo selection for tumorigenic cells had occurred. Two out of 2 mice injected with tumor-derived KMST-G-T24 cells developed 1 cm diameter tumors within 25 days, (a period similar to that of MSU-l-T24 tumorigenic strains), but no decrease in time for tumor development with tumor-derived GM0637B-T24 cells. Both KMST-6-T24 cells and GM0637B-T24 cells expressed T24 H-Las oncogene, but the SUSM-l-T24 cells did not (data not shown), which correlates with their tumorigenicity. 102 DISCUSSION In an earlier study (13) we demonstrated that expression of a transfected T24 H-ras oncogene in normal diploid human fibroblasts resulted in their transformation to morphologically altered cells that form foci and exhibit anchorage-independent growth. Analysis of the relationship between expression of the transfected [as oncogene and tumorigenicity was not possible because expression of the transfected La; oncogene and the transformed phenotype were lost concomitantly during expansion of transformant cell populations. Later, we succesfully isolated several T24 H-ras transformed diploid human fibroblast populations that were stable, i.e., did not revert during expansion in culture subsequent to transfection, and showed that these cells had a finite lifespan in culture and were not tumorigenic (14). In the present study we have extended this research by demonstrating that expression of an exogenous T24 H—m oncogene in a human fibroblast strain that has an infinite lifespan in culture resulted not only in morphological transformation, focus formation and anchorage independence, but also in tumorigenicity. ‘The tumors formed by MSU-l.1-T24 cell strains grew progressively in athymic ice and were diagnosed as invasive fibrosarcomas. It must be noted that cell strains 4 and 5 did not give rise to tumors (Table 1) even though they expressed the T24 H-ras oncogene and exhibited the other transformed characteristics. These 2 strains were retested in a separate experiment and again proved nontumorigenic. The reasons for the lack of tumor formation by these strains are not yet understood. Three other studies have investigated the transforming potential of gas 103 oncogenes in infinite lifespan human fibroblast cell lines. O’Brien et. al. (36) utilized an SV40-virus transformed human fibroblast line, VaZ, and Namba et. al. (37,38) utilized the KMST—6 cell line' .Athough neither of these target cell lines are tumorigenic, they exhibit many characteristics of ras-transformed rodent fibroblasts, i.e., they are morphologically transformed, exhibit aberrant growth behaviour, and are highly aneuploid. Introduction of the r_as oncogene by either virus infection protocols (36,37), or by transfection (38), resulted in cells able to form tumors in athymic mice, although the tumors produced with the VaZ cells did not grow progressively (36). Our results, using cells of the KMST-6 cell line as recipients for transfection of plasmid pH06Tl, confirm the results of Namba et. al. (37,38), and our results using the SV40-transformed infinite lifespan human fibroblast cell line, GM0637B, are similar to those of O’Brien et. al. (36), although our cells formed progressively growing tumors. Because the near-diploid MSU-l.l cell strain exhibits no transformation-related phenotypes, other than its infinite lifespan in culture, it permitted a more complete evaluation of the spectrum of transformation characteristics induced by .:a§_ oncogenes in infinite lifespan human fibroblasts, that are predicted from studies with rodent fibroblasts. Indeed, we feund that expression of the transfected T24 H-Lii oncogene in MSU-1.l fibroblasts induced several transformation-related phenotypes including malignancy. In addition, because of the chromosome stability of MSU-1.1 fibroblasts, we conclude that the transformed phenotypes that resulted following transfection of the T24 H-rgs oncogene was not the result of the selection of tumorigenic 104 variants arising from a karyotypically unstable population of transfectants. Each of the focus-derived MSU-1.1-T24 cell strains generated by transfection of plasmid pH06Tl expressed a high level of T24 H-Las oncogene encoded p21. Transfection of the normal H-ras gene in the same plasmid construct, or the T24 H-ras oncogene in plasmids containing only one or no transcriptional enhancer elements did not result in focus formation. These data suggest that a sufficiently high level of T24 H—ras oncogene expression was necessary to transform MSU-1.1 fibroblasts. Thus, the potent transforming activity of 'the ‘T24 H-[a§_ oncogene appears dependent not only on mutational activation, but also on its enhanced expression. Although the MSU-1.1 cell strain arose following transfection of finite lifespan diploid human fibroblasts with a plasmid containing a v-m oncogene, it is not clear that expression of this oncogene was responsible for induction of the infinite lifespan phenotype since we have only one such cell strain. Likewise, it is not known whether, as proposed for rat embryo fibroblasts (4), the expressed v-myg protein acts in a cooperative fashion with the T24 H-gas oncogene to cause transformation. However, our results showing that ras oncogene expression induces tumorigenic transformation of non-v-myc-transfected infinite lifespan human fibroblasts suggest that the infinite lifespan phenotype, rather than v- myg expression, is the characteristic of MSU-l.1 cells that complemented ras oncogene expression to cause tumorigenicity. Taken together with results showing that finite lifespan human fibroblasts transfected with 105 La; oncogenes exhibit a transformed phenotype, but do not have an infinite lifespan in culture and are not tumorigenic (13,14,19), the results described here indicate both the fundamental and practical importance of the infinite lifespan phenotype for in vitro transformation systems that use tumorigenicity as an endpoint (21). 106 ACKNOWLEDGEMENTS We thank Dr. J. Dillberger for histological examination, Dr. D. Yang for chromosome analysis, Dr. N. Hilkie for kindly supplying us with the pH06 series of plasmids, Dr. M. Anderson for the Yl3-259 antibody to ras p215, Dr. M. Namba for the KHST-6 and SUSM-l human fibroblast cell lines. This research was supported by Department of Energy Grant DEF602-87-ER60524, National Cancer Institute Grant CA21289 and National Institute for Environmental Health Sciences Contract NOI-ES-65152. REFERENCES 1. Andres, A., Schonenberger, C., Groner, B., Hennighausen L., LeMeur, H. & Gerlinger, P. (1987) Pree, Netl, Aged. Sei. CSA 84 1299-1303. 2. Sinn, E., Muller, H., Pattengale, P., Tepler, 1., Wallace, R. & Leder, P. (1987) Cell 49, 465-475. 3. Barbacid, H. (1987) Anne_3eye_fijeeheme 56, 779-8284. 4. Land, H., Parada, L. & Weinberg R. (1983) NngLe_LLengenl 304, 596-602. 5. Ruley, H. E. (1983) Ne;e:e_1Leegenl 304, 602-606. . Parada, L., Land, H, Heinberg, R. , Holf, D. & Rotter, V. (1984) Neture (Lendenl 312, 649- 651. 7. Newbold, R. F. & Overell, R. N. (1983) Neegne_(Lengenl 304, 648-651. 8. Goldfarb, M., Shimizu, K. & Higler, M. Netgre (Lendonl 296, 404-409. 9. Franza, B. R., Maruyama, K., Garrels, J.I. & Ruley, H. E. (1986) Cell 44, 409-418. 10. Spandidos, D. A. 8 Hilkie, N. H. (1984) Netgre (Lengonl 310, 469-475. 11. Pozzatti, R., Muschel, R., Hilliams, J., Padmanabhan, R. Howard, B., Liotta, L. & Khoury, G. (1986) Sejeeee 232, 223- 227. 12. Land, H., Chen, A. C., Morgenstern, J. P., Parada, L. F. & weinberg, R. A. (1986) He], Cell, 819], 6, 1917—1925. 13. Hurlin, P. J., Fry, D. G., Maher, V. H. & McCormick, J. J. Ceneer Res, 47, 5752-5757. 14. Fry, D. G., Hurlin, P. J., Maher, V. M. & McCormick, J. J. Mutat19n_fiess in press. 15. Sager, R. Tanaka, K., Lau, C. C. Ebina, Y. & Anisowic, A. (1983) Proc. Netl. .Acad. Sei , USA 80, 7601- 7605. 16. Spandidos, D. A. (1985) Auticancer Res, 5, 485-498. 17. Feramisco, J. R., Gross, M., Kamata, T., Rosenberg, M. & Sweet, R. H. (1984) Cell 38, 109-117. 18. Stacy, D. N. & Kung, H.-F. (1984) flgCyre (Londenl 310, 508- 107 108 511. 19. Tubo, R. A. & Rheinwald, J. G. (1987) anegene Bee. 1, 407- 421. 20. DiPaolo, J. A. (1983) J. NaLl, Cancer leat, 70, 3-8. 21. McCormick, J. J. & Maher, V. M. (1988) NegationARes. in press. 22. Sach, G. H. (1981) In yigre, 17, 1-19. 23. Namba, M., Nishitani, K., & Kimoto, T. (1978) Jae. J. Exp. Med, 48, 303-311. 24. Namba, M, nishitani, K., Hyodoh, F., Fukushima, F. & Kimoto, T. (1985) Int. J. Caece: 35, 275-280. 25. McCormick, J. J. & Maher, V. M. (1981) in DNA Reeai 1:: abo or Man of Re earch rocedur s eds. Friedberg, EA C. & Hanawalt, P. A. (Marcel Dekker, New York), 501- 522. Bettger, H. J. Boyce, S. T, Halthall, B. J. & Ham, R. G. (1981) Bree, Natl, Ac cad , Sei. USA 78, 5588- 5592. 27. Ryan, P. A., Maher, V. M. & McCormick, J. J. (1987) Exp. Cell Res, 172, 318-328. 28. Southern, P. J. & Berg, P. (1982) J, Mel, Aeel, Cee, 1, 327—341. 29. Morgan, T. L., Maher, V. M. & McCormick, J. J. (1986) 1e_ Vitre Cell. Dev. Biel, 22, 317-319. 30. Heflich, R. H, Dorney, D. J, Maher, V. M. & McCormick, J. J (1977) W 77 634- 641 31. Southern, E. M. (1975) J, Nel, Big]. 98, 503-517. 32. Furth, M. E., Davis, L. J., Fleurdelys, B. & Scolnick, E. M. (1982) Q, Virel, 43, 294-304. 33. Laemmli, U. K. (1970) Nature (Lendonl 227, 680-685. 34. Yunis, J. J. 8 Chandler, M. E. Prog. Clin. Pathol. 7, 267- 288. 35. Schilz, R. J., Maher, V. M. & McCormick, J. J. (1988) J, Cellular Bieehemjstry Supplement 12a, 156. 36. O’Brien, H, Stenman, G. & Sager, R. (1986) Erec. Natl Aeae. Sci. USA 83, 8659-8663. 37. Namba, M., Nishitani, K., Fukushima, F., Kimoto, T. & Nose, 109 K. (1986) Int. 9. Cancer 37, 419-423. 38. Namba, M., Nishitani, K., Fukushima, F., Kimoto, T. & Yuasa, Y. (1988) Ngtatjon Bee, in press. 110 APPENDICIES 111 .czonm mew mmuwm copuuPmemc Amv H mum was sz HHH use: .Amv H: amm_.Axv H mmm .Louum> ooza may cw mcmm mmuuz Fusco: «nu m:_mu:ou Aczosm uozv Hzmoza vwsmm—m .ahoozq use oozq .muwsmMFQ we mam: axn.m_ _hoc:a APPENDIX A .oocv::< m noun: cup 112 APPENDIX B Athymic mouse with tumor formed by subcutaneous injection of MSU-1.1-T24 strain 3 cells. 113 APPENDIX C Data 'not shown“ in Chapter III 114 Figure 1. Comparison of nae p21 expression in focus-derived, anchorage independent colony-derived and tumor-derived MSU-1.1-T24 cell strains. Analysis was performed as described in the Methods and Materials section of Chapter 11. T24 is the T24 bladder carcinoma cell line. 2T, 3T1 and 3T2 are were cultured from independent tumors formed by MSU-l.l-T24 cell strains 2 and 3. 381 and 3B2 are cell strains derived from anchorage independent colonies formed by MSU-1.1-T24 cell strain 3. The bars indicate the location of the normal p21 doublet and the arrow indicates the approximate position of T24 H-[aa p21. 115 116 Figure 2. Southern blot analysis of MC29 sequences in MSU-l.1-T24 cell strains. Genomic DNAs were digested with XhQ I, and Cam, HI, and hybridized with the Sag I fragments of the MC29 virus (Reddy et al. (1983) Proc. Natl. Acad. Sci. USA CC, 2500-2504). Hybridizing sequences specific to MC29 viral sequences are indicated. 117 H 7' a a MSU—LI-T24 Strains *3 E 1 2 6 23- 9.4- 6.6- 4.4- -> v- s“ — 118 Figure 3. Analysis of pa; p21 expression in SUSM-l-T24, KMST-6-T24 and GM0637-T24 cell strains and in their parent cell lines. Analysis was performed as described in Chapter II Maaterials and Methods with the exception that immunoprecipitates were solubilized in SDS-PAGE buffer containing 8 M urea. This modification caused a shift in the mobility of the mutant p21 so that it migrated to a position above the characteristic p21 doublet (see arrow). T24 is the human bladder carcinoma cell line. The bars indicate the location of the normal p21 doublet and the arrow indicates the location of the T24 H-Laa encoded p21. ‘_—r—*‘_ 119 vwfiipmmczw panes—U vufilmufimzx wufimEM vmblfiuzwbm HIEmDm «2.. H04