AV'O-JE‘SAS VEIIBRSTYL IIIIIIIIIIIIIIIIIIIIIIIIIIII I II IIIIIZIIIIIIIII 3 1293 015702 This is to certify that the dissertation entitled ‘ ' ' JIKC‘h'Ob) lflV€5+jaton 0% (alfaum Lj/lfil 'f/flhS Faf‘dfijfl) 4;” Auw“ J",-0£/&Sf presented by SAW“ IL/Mj has been accepted towards fulfillment of the requirements for fi' D degree in 67’497‘"; \\QJI ‘W G) Major professor Date MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State Unlverslty PLACE ll RETURN BOX to remove thle checkout from your record. To AVOID FINES return on or betore date due. DATE DUE DATE DUE DATE DUE MSU le An AffirmiLve Action/Equal Opportunity lnetltuton W M! INVESTIGATION OF Ca“ SIGNAL TRANSDUCTION PATHWAY(S) IN HUMAN FIBROBLASTS By Shixia Huang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Program 1997 ABSTRACT INVESTIGATION OF Ca“ SIGNAL TRANSDUCTION PATHWAY(S) IN HUMAN FIBROBLASTS By Shixia Huang In serum-free medium containing serum replacements but totally lacking in protein growth factors, diploid human fibroblasts remained quiescent if the extracellular Ca2+ concentration was only 0.1 mM. However, when the Ca2+ concentration in this medium was increased to 1 mM, the cells replicated as rapidly as they do in medium supplemented with protein growth factors. When quiescent cells in medium with only 0.1 mM Ca2+ were exposed to 1 mM or 10 mM Ca2+ or 100 ng/ml EGF, the 42 kDa and 44 kDa forms of MAPK were rapidly activated, as demonstrated by a characteristic electrophoretic mobility shift of these proteins and by their enhanced ability to phosphorylate myelin basic protein (MBP). Analysis of fractions from Mono Q anion-exchange chromatography of lysates of cells exposed to 10 mM Ca2+ or 100 ng/ml EGF revealed a peak of MBP phosphorylation activity that coeluted with p42 and p44 MAPK as shown by immunoblot analysis. Activation of MAPK by extracellular Ca2+ was dose dependent and biphasic, with a peak of activation at 5-10 min afier exposure, followed by a period of sustained activation of MAPK at a lower level. This pattern has been shown (Vouret-Craviari et al. [1993], Biochem J. 289, 209) to correlate with the re-entry of mammalian cells into the cell cycle. We then investigated the pathway of Cap-induced MAPK activation. We found the exposure of human fibroblasts to extracellular Ca2+ increased cytosolic inositol (l, 4, 5)- trisphosphate level and caused a transient rise in the concentration of intracellular Ca2+. Ca2+- induced MAPK activation was partially abolished by treatment of the cells with pertussis toxin which inhibits certain species of G proteins. It was also decreased by treatment with thapsigargin which depletes intracellular Ca2+ stores, by exposure to phorbol 12—myristyl 13- acetate that downregulates protein kinase C (PKC), by treatment with calmodulin antagonists, N-(6-aminohexyl)-5-chloro—l~naphthalenesulfonamide HCL and calmidazolium, as well as by exposure of the cells to lanthanum, a Ca“ channel inhibitor. We found that the c-raf-l protein was not phosphorylated after Ca” stimulation, and the Ca2+ sensing receptor (Brown et al. [1993] Nature 366, 575-580) was not involved in mediating MAPK activation or cell growth. These results suggest that extracellular Ca“ stimulates MAPK activation through a pathway(s) involving a pertussis toxin sensitive G protein, phospholipase C, intracellular Ca2+, calmodulin, and PKC. This work is dedicated to my husband: Yi Li my daughter: Grace (DanDan) my unborn baby my father: J ia Gu Huang my mother: Dong Xiu Li ACKNOWLEDGMENTS I am grateful to my graduate advisor, Dr. J. Justin McCormick, for his guidance, his encouragement, and his support throughout my graduate program. I also want to express my deep appreciation to Dr. Veronica Maher for her advice, her support, and her selfless efforts in revising my papers. I would like to give special thanks to my other committee members: Dr. Susan Conrad, Dr. David DeWitt, and Dr. Walter Esselman, for their advice, their encouragement, and their precious time. I would like to thank Dr. Joseph Przybyszewski for his help in editing Chapter I. I would also like to thank members of the Carcinogenesis Laboratory for their help and friendship during my graduate years. Finally, I am thankful for the support and help of my husband Yi. I want to thank my daughter Grace (DanDan) who provides me with so much love and fun and makes my life so enjoyable, and my unborn baby whose cooperation has made the writing of my dissertation possible in a relatively short time. Special thanks also go to my parents whose love and encouragement are very important to me. TABLE OF CONTENTS LIST OF FIGURES ............................................................................................................. ix ABBREVIATIONS ............................................................................................................. x INTRODUCTIONS ............................................................................................................. 1 List of references ..................................................................................................... 4 CHAPTER 1. LITERATURE REVIEW ............................................................................... 7 A. Extracellular Ca2+ as the first messenger ............................................................ 7 l. Extracellular Ca2+ and cell growth .......................................................... 7 2. Other functions of extracellular Ca2+ ..................................................... 10 B. Growth factor pathways 1. Growth factors and their receptors ......................................................... 11 2. Intracellular pathways 2.1 Overview of the pathways upon growth factor stimulation ............ 13 2.2 G proteins ...................................................................................... .18 2.3 Phospholipase C ............................................................................. .19 2.4 Protein kinase C ............................................................................. 22 2.5 Raf ................................................................................................. .24 2.6 Intracellular Ca2+ ............................................................................. 27 2.7 Calmodulin .................................................................................... .29 2.8 MAP kinase ................................................................................... .33 C. Ca2+ signaling in mammalian cells 1. Intracellular components involved in Ca2+ signaling and cell differentiation/hormonal secretion ................................................... 35 2. Intracellular components involved in Ca2+ stimulated growth pathway..37 3. Ca2+-sensing receptor and its signaling .................................................. 38 4. Other possible Cay-sensing receptors ..................................................... 39 List of references .................................................................................................... 42 CHAPTER II. Extracellular Ca2+ stimulates the activation of mitogen-activated protein kinase and cell growth in human fibroblasts ................................. 65 Abstract .................................................................................................................. 66 Abbreviations ......................................................................................................... 67 Introduction ................................................................................... 68 vi Materials and methods Materials ..................................................................................................... .70 Cells and cell culture .................................................................................. 70 Media for testing for growth stimulation ................................................... 70 Preparation of cell lysates ........................................................................... 71 Immunoblot analysis of MAPK ................................................................. 72 MBP phosphorylation as an assay for MAPK activity .............................. 72 Purification of MAPK by Mono Q column chromatography ..................... 73 Results and discussion Evidence that extracellular Ca2+ stimulates cell growth ............................ .74 MAPK activation as a result of Ca” or EGF stimulation .......................... 74 Time course for Cay-induced MAPK activation ...................................... .81 Acknowledgments ................................................................................................... 85 References ............................................................................................................. .86 CHAPTER III. Characterization of the second messenger pathway of extracellular Ca“- induced mitogen-activated protein kinase activation in human fibroblasts ................................................................................................................ 88 Abstract .................................................................................................................. 89 Abbreviations ................................................................................. 90 Introduction ........................................................................................................... .91 Material and methods Materials ..................................................................................................... .94 Cells and cell culture ................................................................................. .94 Isolation and culture of mouse skin fibroblasts ......................................... 95 Media for testing for growth stimulation of mouse fibroblasts .................. 96 Preparation of cell lysates .......................................................................... 96 Immunoblot analysis of MAPK (Western Blot) ......................................... 96 MAP-2 phosphorylation assay (MAPK activity assay) .............................. 97 Intracellular Ca2+ measurement ................................................................. .97 Cytosolic 1P3 measurement ....................................................................... 98 Results and discussion Evidence that activation of MAPK by extracellular Ca2+ is partially inhibited by PMA ............................................................ 100 Evidence that extracellular Ca” c-raf-l phosphorylation .......................... 102 Evidence that extracellular Ca2+ induces intracellular Ca2+ mobilization .................................................................................. 104 Evidence that the mobilization of intracellular Ca2+ transient by extracellular Ca” is partially abolished by thapsigargin ................................... 106 Evidence that activation of MAPK by extracellular Ca2+ is partially inhibited by thapsigargin ............................................................. 107 Evidence that activation of MAPK by extracellular Ca” is abolished by the channel blocker lanthanum, but not by verapamil or nifedipine ................................................................................. 109 Evidence that activation of MAPK by extracellular Ca2+ is partially vii abolished by calmodulin antagonists ............................................ 112 The role of PLC ........................................................................................ 114 Evidence that MAPK activation by extracellular Ca2+ is partially abolished by pertussis toxin .......................................................... 118 Evidence that the Cay-sensing receptor is not involved in extracellular Ca”- induced MAPK activation in mouse skin fibroblasts ................... .120 References ............................................................................................................ . 125 viii LIST OF FIGURES CHAPTER 11 Figure 1. Growth stimulating activity of Ca2+ ................................................................... .75 Figure 2. MAPK activation as a result of Ca2+ or EGF stimulation .................................. .77 Figure 3. Assay of Mono Q fractions for ability to phosphorylate MBP .......................... .79 Figure 4. Immunoblot analysis of Mono Q fractions ....................................................... .80 Figure 5. Dose-response of Cay—stimulated MAPK activity ............................................ .82 Figure 6. Time course for Ca2+-stimulated MAPK activity ................................................ 84 CHAPTER III Figure 1. Activation of MAPK by extracellular Ca2+ is partially inhibited by PMA ....... 101 Figure 2. Extracellular Ca2+ does not induce c-raf-l phosphorylation ............................ 103 Figure 3. Intracellular Ca2+ mobilization after Ca2+ stimulation ...................................... 105 Figure 4. Effect of thapsigargin, or PMA and thapsigargin together, on Ca2+- or EGF-induced MAPK activation .................................................................... 108 Figure 5. Effect of Ca2+ channel blockers on Ca2+ or EGF-induced MAPK activation...110 Figure 6. Effect of calmodulin antagonist on Ca2+ or EGF-induced MAPK activation...1 13 Figure 7. Cytosolic level of 1P3 was increased after Ca2+ stimulation. ........................... 116 Figure 8 Effect of extracellular Ca2+ on PLCy phosphorylation ..................................... 117 Figure 9. Activation of MAPK by extracellular Ca2+ is partially inhibited by pertussis toxin ............................................................................................................. 119 Figure 10. MAPK activation by extracellular Ca2+ and PBS in fibroblasts from mice that differ in Cay-sensing receptor capability ........................................................................ 121 BSA CaM l CaMK CAMP cdks CR1 CR2 CR3 DAG EGF EGFR ER FBS F GF 3 GAP GDP GTP G proteins lGFs 1P3 1P4 ABBREVIATIONS bovine serum albumin calmodulin l calmodulin dependent protein kinase cyclic adenosine 3', 5'-phosphate cyclin dependent kinases conserved region 1 in raf kinase conserved region 2 in raf kinase conserved region 3 in raf kinase diacylglycerol epidermal growth factor EGF receptor endoplasmic reticulum fetal bovine serum fibroblast growth factors GTPase activating protein guanosine 5'-diphosphate guanosine 5'-triphosphate guanine nucleotide binding proteins insulin-like growth factors inositol (1, 4, 5)-tn'sphosphate inositol (l , 3, 4, 5)—tetrakisphosphate IRS JAK MAP-2 MAPK MBP MEK NGF NMDA PDGF PI3 kinase PIP2 PLC PKC PMA PTH SDS-PAGE SH2 SH3 STAT TGFB W-7 W-12 insulin receptor substrate Janus kinase microtubule-associated protein 2 mitogen-activated protein kinase myelin basic protein MAPK kinase nerve growth factor N-Methyl-D-aspartate platelet-derived growth factor phosphotidylinosite 3 kinase phosphatidylinositol 4, 5-bisphosphate phospholipase C protein kinase C phorbol 12-myristyl 13-acctate parathyroid hormone sodium dodecyl sulphate-polyacrylamide gel electrophoresis src homology 2 src homology 3 signal transducer and activator of transcription tumor growth factor [3 N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide HCL N-(4-aminobutyl)-2-naphthalenesulfonamide xi INTRODUCTION It has been known since the 1970's that Ca2+ is required for normal mammalian cells to grow in culture (1-4). The first evidence of Ca2+ acting as a growth promoter came from a study with mouse 3T3 fibroblasts by Dulbecco and Belkington (5), in which they showed that increasing the concentration of Ca2+ above the physiological level (1.8 mM) promoted DNA synthesis. Previous colleagues (6,7) in Carcinogenesis Laboratory at Michigan State University have shown that human fibroblasts grow in a serum free medium supplemented with 1.0 mM Ca2+ as rapidly as in this medium with serum or growth factors when supplemented with the serum replacements, and in the presence of insulin as the only protein growth factor. Although normal fibroblast cells require 1-2 mM Call for their growth, studies (6-8) have shown that fibroblasts transformed in culture or derived from fibrosarcomas are able to replicate in culture in serum free medium containing 0.1 mM Ca2+ or at a lower concentration, but no protein growth factors. This indicates that 1 these cells have some alteration in the pathway by which Ca2+ drives growth. Since the mechanism of Cay-induced cell growth is not understood, thus studying the mechanism of Cay-induced cell growth may provide insight into the transformation process. Therefore, my studies concentrated on the elucidation of the mechanism(s) of Cali-driven growth in normal human fibroblasts. It is well known that protein growth factors,, such as epidermal growth factor (EGF), platelet-derived growth factor (PDGF), insulin, thrombin, nerve growth factor, and bombesin, bind to specific receptors on the external cell surface, and initiate phosphorylation cascades through the activation of either tyrosine kinase-linked or guanine nucleotide l 2 binding protein (G protein)-coupled receptors, and one of the key events in these cascades is the activation of mitogen-activated protein kinase (MAPK) by phosphorylation (9-15). Several closely related MAPK isoforms, referred to by their molecular masses, i.e., p40, p42, p44, and p54 (16), have been identified in various cell types. In human fibroblasts, it has been shown that EGF (11), tumor necrosis factor (13,17,18) and interleukin (18) can cause activation of p42 and/or p44 MAPKs. The activation of MAPK occurs when activated MAPK kinase phosphorylates the tyrosine and serine/threonine residues of MAPK (19). This phosphorylation can be detected because it results in a mobility shift of MAPK in sodium dodecyl sulphate—polyacrylamide gel electrophoresis (SDS-PAGE). The activated MAPK can, in turn, phosphorylate microtubule-associated protein (MAP-2) (20), myelin basic protein (MBP) (21), ribosomal S6 kinase (22), the EGF receptor (23), phospholipase A2 (24), and the nuclear transcriptional factors c-myc (25,26), c-fos, and c-jun (23,27). In mouse 3T3 fibroblasts in culture, it has been shown that extracelluar Ca2+ stimulates DNA synthesis (28-33) and c-fos expression (30-32,34) at a level similar to that induced by PDGF (31,32). Furthermore, Epstein et al. (31,32) showed that both PDGF- induced and Ca2 +-induced c-fos expression is dependent on protein kinase C (PKC). However, unlike PDGF, Ca2+ did not induce the phosphorylation of the PDGF receptor or the raf protein, indicating that the pathways which result in similar gene transcription are, at least in part, independent. The objectives of the present work were (1) to demonstrate that human fibroblasts can grow in a medium containing 1.0 mM Ca2+ and serum replacements but no protein growth factors and to determine whether Ca2+ stimulates the activation of MAPK; (2) to investigate the involvement of various intracellular components in the pathway of Ca2+- 3 induced MAPK activation; (3) to investigate the involvement of certain membrane elements, such as the Cay-sensing receptor of Brown et al. (35) in the pathway of Cay-induced MAPK activation and cell growth. Chapter 1 reviews the literature regarding extracellular Ca2+ as the first messenger, growth factor pathways, and Ca2+ signaling in mammalian cells. Chapter II consists of a manuscript published in the Biochemical Journal , 310, 881-885, 1995. It describes the research I carried which demonstrates that in normal human fibroblasts extracellular Ca” can stimulate the activation of MAPK and cell growth in the absence of protein growth factors, and correlates the MAPK activation with the growth of these cells under such conditions; Chapter III consists of a manuscript to be submitted to the Biochemical Journal. It describes the research I carried which demonstrates the pathways involved in extracellular Cap-induced MAPK activation. 4 LIST OF REFERENCES l. Hazelton, B., Mitchell, B., and Tupper, J. Calcium, magnesium, and growth control in the WI-38 human fibroblast cell. J. Cell Biol., 83: 487-498, 1979. 2. Balk, S.D., Polimeni, P.I., Hoon, B.S., LeStourgeon, D.N., and Mitchell, R.S. Proliferation of Rous sarcoma virus-infected, but not of normal, chicken fibroblasts in a medium of reduced calcium and magnesium concentration. Proc. Natl. Acad. Sci. U. S. A., 76: 3913-3916, 1979. 3. Balk, S.D., Whitfield, J.F., Youdale, T., and Braun, A.C. Roles of calcium, serum, plasma, and folic acid in the control of proliferation of normal and Rous sarcoma virus-infected chicken fibroblasts. Proc. Natl. Acad. Sci. U. S. A., 70: 675-679, 1973. 4. Balk, S.D. Calcium as a regulator of the proliferation of normal, but not of transformed, chicken fibroblasts in a plasma-containing medium. Proc. Natl. Acad. Sci. U. S. A., 68: 271-275, 1971. 5. Dulbecco, R. and Elkington, J. Induction of growth in resting fibroblastic cell cultures by Ca++. Proc. Natl. Acad. Sci. U. S. A., 72: 1584-1588, 1975. 6. Wilson, D.M., Yang, D.J., Dillberger, J.E., Dietrich, S.E., Maher, V.M., and McCormick, J.J. Malignant transformation of human fibroblasts by a transfected N-ras oncogene. Cancer. Res., 50: 5587-5593, 1990. 7. Morgan, T.L., Yang, D.J., Fry, D.G., Hurlin, P.J., Kohler, S.K., Maher, V.M., and McCormick, J.J. Characteristics of an infinite life span diploid human fibroblast cell strain and a near-diploid strain arising from a clone of cells expressing a transfected v-myc oncogene. Exp. Cell Res., 197: 125-136, 1991. 8. SchilL R.J. Expression of growth factor genes in transformed human fibroblasts and cells derived from human fibrosarcomas: a possible mechanism for replication in the absence of exogenous growth factors. Ph.D Dissertation. East Lansing: Michigan State University, 1 988. 9. Pelech, S.L. and Sanghera, J.S. MAP kinases: charting the regulatory pathways [comment]. Science, 257: 1355-1356, 1992. 10. Burgering, B.M., de Vries Smits, A.M., Medema, R.H., van Weeren, P.C., Tertoolen, LG, and Bos, J .L. Epidermal growth factor induces phosphorylation of extracellular signal-regulated kinase 2 via multiple pathways. . Mol. Cell Biol., 13: 7248-7256, 1993. 11. Chao, T.S., Byron, K.L., Lee, K.M., Villereal, M., and Rosner, M.R. Activation of MAP 5 kinases by calcium-dependent and calcium-independent pathways. Stimulation by thapsigargin and epidermal growth factor. J. Biol. Chem., 267: 19876-19883, 1992. 12. Nakielny, S., Cohen, P., Wu, J., and Sturgill, T. MAP kinase activator from insulin-stimulated skeletal muscle is a protein threonine/tyrosine kinase. EMBO. J., 11: 2123-2129, 1992. 13. Victor, 1., Schwenger, R, Li, W., Schlessinger, J., and Vilcek, J. Tumor necrosis factor-induced activation and increased tyrosine phosphorylation of mitogen-activated protein (MAP) kinase in human fibroblasts. J. Biol. Chem, 268: 18994-18999, 1993. 14. Pang, L., Decker, S.J., and Saltiel, AR Bombesin and epidermal growth factor stimulate the mitogen-activated protein kinase through different pathways in Swiss 3T3 cells. Biochem. J., 289: 283-287, 1993. 15. Lenorrnand, P., Sardet, C., Pages, G., L'Allemain, G., Brunet, A., and Pouyssegur, J. Growth factors induce nuclear translocation of MAP kinases (p42mapk and p44mapk) but not of their activator MAP kinase kinase (p45mapkk) in fibroblasts. J. Cell Biol., 122: 1079-1088, 1993. 16. Pelech, S.L. and Sanghera, J .S. Mitogen-activated protein kinases: versatile transducers for cell signaling. Trends. Biochem. Sci., 1 7: 233-23 8, 1992. 17. Victor, 1. and Vilcek, J. Pathways of heat shock protein 28 phosphorylation by TNF in human fibroblasts. . Lymphokine. Cytokine. Res., 13: 315-323, 1994. 18. Guy, G.R., Chua, S.P., Wong, N.S., Ng, SB, and Tan, Y.H. Interleukin 1 and tumor necrosis factor activate common multiple protein kinases in human fibroblasts. J. Biol. Chem, 266: 14343-14352, 1991. 19. Wu, J., Michel, H., Dent, P., Haystead, T., Hunt, DR, and Sturgill, T.W. Activation of MAP kinase by a dual specificity Tyr/Thr kinase. Adv. Second. Messenger. Phosphoprotein. Res., 28: 219-225, 1993. 20. Sillirnan, CC. and Sturgill, T.W. Phosphorylation of microtubule-associated protein 2 by MAP kinase primarily involves the projection domain. Biochem. Biophys. Res. Commun., 160: 993-998, 1989. 21. Sanghera, J.S., Aebersold, R., Monison, H.D., Bures, E.J., and Pelech, S.L. Identification of the sites in myelin basic protein that are phosphorylated by meiosis-activated protein kinase p44mpk. FEBS. Lett., 273: 223-226, 1990. 22. Sturgill, T.W., Ray, L.B., Erikson, E., and Maller, J .L. Insulin-stirnulated MAP-2 kinase phosphorylates and activates ribosomal protein S6 kinase 11. Nature, 334: 715-718, 1988. 6 23. Davis, R.J. The mitogen-activated protein kinase signal transduction pathway. J. Biol. Chem, 268: 14553-14556, 1993. 24. Lin, L.L., Wartmann, M., Lin, A.Y., Knopf, J.L., Seth, A., and Davis, R.J. cPLA2 is phosphorylated and activated by MAP kinase. Cell, 72: 269-278, 1993. 25. Gupta, S. and Davis, R.J. MAP kinase binds to the NH2-terminal activation domain of c-Myc. FEBS Lett., 353: 281-285, 1994. 26. Seth, A., Gonzalez, F.A., Gupta, S., Raden, D.L., and Davis, R.J. Signal transduction within the nucleus by mitogen-activated protein kinase. J. Biol. Chem, 26 7: 24796-24804, 1992. 27. Davis, R.J. Transcriptional regulation by MAP kinases. Mol. Reprod. Dev., 42: 459-467, 1 995 . 28. Rubin, H. and Sanui, H. Complexes of inorganic pyrophosphate, orthophosphate, and calcium as stimulants of 3T3 cell multiplication. Proc. Natl. Acad. Sci. U. S. A., 74: 5026-5030, 1977. 29. Bowen Pope, DP. and Rubin, H. Growth stimulatory precipitates of Ca2+ and pyrophosphate. . J. Cell Physiol., 11 7: 51-61 , 1983. 30. Mitchell, P.G., Pledger, W.J., and Cheung, H.S. Molecular mechanism of basic calcium phosphate crystal-induced mitogenesis. Role of protein kinase C. J. Biol. Chem, 264: 14071-14077, 1989. 31. Epstein, R.J., Druker, B.J., Irminger, J .C., Jones, S.D., Roberts, T.M., and Stiles, C.D. Extracellular calcium mimics the actions of platelet-derived growth factor on mouse fibroblasts. Cell Growth Differ, 3: 157-164, 1992. 32. Epstein, R. Calcium-inducible transmodulation of receptor tyrosine kinase activity. Cell Signal., 7: 377-388, 1995. 33. Hennings, H., Michael, D., Cheng, C., Steinert, P., Holbrook, K., and Yuspa, S.H. Calcium regulation of growth and differentiation of mouse epidermal cells in culture. Cell, 19: 245-254, 1980. 34. Cheung, H.S., Mitchell, P.G., and Pledger, W.J. Induction of expression of c-fos and c-myc protooncogenes by basic calcium phosphate crystal: effect of beta-interferon. Cancer Res., 49: 134-138, 1989. 35. Brown, B.M., Gamba, G., Riccardi, D., Lombardi, M., Butters, R., Kifor, 0., Sun, A., Hediger, M.A., Lytton, J., and Hebert, S.C. Cloning and characterization of an extracellular Ca(2+)-sensing receptor fi'om bovine parathyroid. Nature, 366: 575-580, 1993. CHAPTER I LITERATURE REVIEW A. Extracellular Ca2+ as the first messenger l. Extracellular Ca2+ and cell growth In normal resting mammalian cells, the cytosolic free Ca2+ concentration is maintained at 10-100 nM (1), in contrast, the extracellular free Ca2+ concentration in blood and tissue culture medium, is between 1-2 mM. “The importance of intracellular Ca” in regulating cell function is well recognized. Not less important, but less well understood, is the fundamental role played by extracellular Ca2+ in the modulation of cell function ” (2). Studies in the 19703’ have shown that there is a strict requirement of 1-2 mM extracellular Ca2+ (physiological level) for the growth of chicken (3-5), mouse (6), and human (7) fibroblasts in culture. However, the mechanism(s) which set this Ca2+ requirement for normal fibroblasts to grow in culture is not clear yet. A study by Dulbecco and Elkington (6) in 1975 showed that when added to resting cultures of mouse 3T3 fibroblasts in serum-free medium, extracellular Ca2+ concentrations above 1.8 mM induced DNA synthesis and cell growth. Additionally, they showed that the growth rate in 5.4 mM Ca2+ was similar to that observed after the addition of 10% serum. They also found that the effect of Ca2+ on cell growth was synergistic with that of serum This was the first report showing that Ca2+ itself acted as a growth promoter. Several reports (8-15) support and expand these findings and suggest some insight into the mechanisms by which Ca2+-induced cell growth in 3T3 mouse fibroblasts. These will be discussed in detail below. 8 Hazelton et al. (7) showed that physiological levels of extracellular Ca2+ were required for human lung fibroblasts (WI-3 8) to enter 8 phase. When Ca” in the medium was reduced to 0.06 mM or lower, a substantial fraction of WI-38 cells were prevented from entering S phase. When the Ca2+ concentration in the medium was increased to 2 mM, the cells entered S phase. Their study also showed that, in contrast to normal WI-38 cells, SV- 40 transformed WI-38 cells did not require Ca2+ to enter S phase. Since then, another group (16) showed similar Ca2+ requirements (1-2 mM) for the growth of human lung fibroblasts (WI-38 and IMR90) in medium containing serum pretreated with chelex to remove the Ca2+, but not for that of their simian virus 40 (SV40) transformed counterparts. In 1986, Praeger and Cristofalo (17,18) showed that WI-38 cells can grow in a serum free medium with Ca2+ (5 mM) or epidermal growth factor (EGF) in the presence of insulin, dexamethasone and transferrin. They also showed that combining suboptimal doses of EGF and Ca2+ resulted in an additive effect on cell growth. In 1977, Pledger et. al. (19) showed that serum contains two types of growth promoting agents: competence factors, e. g., platelet-derived growth factor (PDGF), and progression factors, e. g., insulin-like growth factors (IGFs). These factors function synergistically to promote growth of BALB/c 3T3 mouse fibroblasts (19,20). Quiescent 3T3 cells exposed briefly to PDGF become “competent” to replicate their DNA and divide; however, these “competent” cells did not progress efficiently into S phase unless they were incubated throughout Go/G, with medium containing plasma which contains progression factor (19). Studies by Rubin and Samui (9) and Bowen Pope and Rubin (10) suggest that Ca2+ stimulated cell growth by forming precipitates with inorganic pyrophosphate or orthophosphate. In their studies, they observed that Ca2+-stimulated DNA synthesis 9 correlated with the concentration of inorganic orthophosphate or pyrophosphate in the medium, as well as the turbidity of the medium. They proposed that precipitates of Ca2+- phosphate complexes acted in a manner similar to macromolecular stimulants such as hormones and interacted with the plasma meinbrane. In 1979, Stiles et al. (20) found that Ca3(PO4)2, like PDGF, acted early in the cell cycle as a competence factor, but did not act as a progression factor. Mitchell et al. (11) and Cheung et al. (12) found that basic calcium phosphate crystals, like PDGF, induced c-fos and c-myc gene expression. These above studies were carried out in mouse 3T3 fibroblasts using DME medium with 5% platelet poor plasma which contained the progression factors necessary for mouse fibroblasts to enter the cell cycle after stimulation by competence growth factors, such as PDGF. The distinct role of competence and progression factors has been demonstrated in mouse 3T3 fibroblasts (19,20) and lymphocytes (21). Human fibroblasts appear to need only competence growth factor such as PDGF or EGF for their growth. One explanation for this is that human fibroblasts synthesize IGF-l (22) during cell proliferation, and this IGF-l may act as a progression factor. It is also known that during proliferation, human fibroblasts secrete fibroblast growth factor (F GF), and FGF itself is a mitogen for fibroblasts (23). It has been suggested that FGF may act in a positive feedback loop to induce proliferation (23). Sugimoto et al. (24) showed that in osteoblasts, cell growth is stimulated by high concentration of Ca2+, and that was mediated by IGF -1. In the Carcinogenesis Laboratory of Michigan State University, Morgan et al. (25) showed that diploid human foreskin fibroblasts can replicate rapidly in serum-flee McM medium, a modified version of MCDB 110 (26), if supplied with designated serum replacements and with insulin as the only protein growth factor, instead of insulin and EGF 10 as described by Ryan et al. (27). In this medium, supplemented with serum replacements factors normal fibroblasts could grow for their entire in vitro life time (McCormick et al., unpublished results). The McM medium used in these studies contained 1.0 mM Ca2+ (25,28). Morgan et al. (25) also showed that normal human fibroblasts did not replicate if the Ca2+ concentration was reduced from 1 mM to 0.1 mM, suggesting that extracellular Ca2+ stimulates cell growth. 2. Extracellular Ca2+ and other functions In addition to its involvement in cell proliferation, extracellular Ca2+ has also been shown to function in hormonal secretion (29) and cell differentiation (30-35). Studies (30,32-34) showed that mouse and human keratinocytes in culture proliferated in a medium with a low Ca2+ concentration (0.05-0.1 mM), and when the Ca2+ concentration was increased to 1.4-2.0 mM, these cells underwent terminal differentiation. In parathyroid cells, the decrease of extracellular Ca2+ induces the secretion of parathyroid hormone (PTH) (29,36). Extracellular Ca2+ also negatively regulates the growth of parathyroid cells (36). Sakaguchi (36) found that the acidic fibroblast growth factor autocrine system was a mediator of Ca2+-regulated parathyroid cell growth. Expression of both the mRNA and peptide of acidic F GF was suppressed by increasing the extracellular Ca” concentration. Kidney cells, osteoclasts, and placental cells also have the ability to sense changes in the extracellular Ca2+ concentration (29). Osteoclasts, the cells that resorb bone during the skeletal remodeling cycle, recognize and respond to changes in the extracellular Ca2+ concentration in a manner similar to that of parathyroid cells. Increasing the concentration ll of extracellular Ca2+ reduces the expression of podosomes, which contribute to cellular attachment to bone, and also decreases their bone resorptive activity (29,37). In placental cells, high concentrations of extracellular Ca2+ inhibit the secretion of the polypeptide hormone, placental lactogen, from placental fragments (38,39) as well as from dispersed or enriched cultures of trophoblastic cells prepared from placenta. B. Growth factor pathways 1. Growth factors and their receptors Cells exchange growth and differentiation signals through grth factors and hormones (40,41). Most growth factors and hormones are unable to pass the hydrophobic cell membrane. Instead, they exert their effects by binding to specific receptor proteins on the cell surface of the target cells (41). These cell surface receptors bind to the signal molecule (e.g., a protein growth factor) and convert the extracellular signal into intracellular events that affect the behavior of target cells. Cell surface receptor proteins are divided into three classes: enzyme-linked receptors, GTP binding protein (G protein)-coupled receptors, and ion channel-linked receptors (42). Ion channel-linked receptors, which are also known as transmitter-gated ion channels, mediate their signals by neurotransmitters that are involved in rapid synaptic signaling between electrically excited cells. Growth factors or hormones, which promote cell growth in their target cells, exert their effects through the activation of either enzyme-linked or G protein-coupled receptors (43). With enzyme-linked receptors, the extracellular binding of a ligand causes the 12 activation of receptors that function directly as enzymes, such as tyrosine kinases, or are associated with enzymes (42,44). Most of the receptors in this family have an extracellular domain with ligand binding sites, a single transmembrane domain, and an intracellular domain with catalytic sites (41,42). Growth factors that have this kind of receptor include EGF, PDGF, FGF, and IGF, etc. (41). The majority of receptors of this type are protein kinases, or are associated with protein kinases, each of which phosphorylates specific sets of proteins in a target cell. Most of protein tyrosine kinase receptors undergo dimerization upon ligand binding (41). After dimerization, these receptors undergo autophosphorylation, and in turn, phosphorylate downstream kinases which are associated with the receptor molecules by their src homology 2 (SH2) domain (45,46). G protein-coupled receptors bind to their ligands, and transmit a signal through a group of G proteins, which in turn regulate downstream effectors, such as enzymes and ion channels (42,44). All G protein-coupled receptors have an extracellular domain which binds to the ligand, a common seven transmembrane domain, which is structurally. distinct from the single transmembrane domain of the enzyme-linked receptors, and an intracellular domain which interacts with G proteins (42). Growth factors or hormones which bind to this kind of receptor include bradykinin, thrombin, bombesin, interleukin 8, parathyroid hormone, somatostatin, A2 adenosine, etc. (29,44,47). The G protein-coupled receptors respond to the binding of a specific ligand by undergoing a conformational change. This allows the intracellular domain of the receptor to interact with a specific type(s) of G proteins. If a receptor can interact with only one subtype of G protein, it will activate effectors only through this G protein, and the response will be relatively focused. On the other hand, if a receptor interacts with several G-proteins, each of the G proteins will interact 13 with one or more effectors, and the response will be more diffirse since it will involve several different pathways. The function of G proteins will be discussed in section 2.2. 2. Intracellular pathways 2.] Overview of the pathways upon growth factor stimulation Growth factors or hormones promote cell proliferation by binding and activating their specific receptors, mainly through either tyrosine kinase-linked receptors, or G protein- coupled receptors, and the activation of these receptors initiates one or more phosphorylation cascade(s) (48-54) (see figure on next page and detail description below). Pathways involving the tyrosine kinase-linked receptors The activation of receptor tyrosine kinases activates downstream effectors. As described above, the binding of a ligand to a receptor induces dimerization and autophosphorylation of the receptor . The individual phosphotyrosine residues located in the cytoplasmic domains of receptors serve as highly selected binding sites that interact with specific cytoplasmic molecules (55). A common feature of many signaling molecules that are known to bind to receptor tyrosine kinases with high affinity is that they contain src homology 2 (SH2) domains, regions of about 100 amino acids that have homology with the noncatalytic region present in the c-src proto-oncoprotein (56-58). A large number of proteins containing SH2 domains, and the specific sequences that these proteins recognize and bind to, have been identified (57,58). The SH2 domain of the p85 subunit of phosphoinositide 3 kinase (PI3 kinase) has been shown to associate with the human PDGF [3 receptor via tyrosines 740 and 751, which lie in Growth Factor Pathways ~.~——y—. <§ene Transcription 15 the kinase insert regions of this receptor (59), with the murine colony-stimulating factor-1 (CSF-l) receptor via tyrosine 721 (60), and with several phosphorylated tyrosines on the insulin receptor substrate-1 (IRS-1) (61). All of these tyrosine residues lie within the consensus sequence of tyrosine x x methionine (YxxM) (where x can be a wide range of possible residues), which is important for PI3 kinase binding (62). Other SH2 domain- binding sites have been identified on the human PDGF B-receptor. These include tyrosine 771 in the kinase insert region, the site for interaction with ras GTPase activating protein (rasGAP) (63), tyrosine 1020 in the C-terminal tail for interaction with phospholipase C y (PLCy) (64), tyrosine 1009, the binding site for the phosphotyrosine phosphatase, syp (65), and tyrosine 579 and 581 in the juxarnembrane segment N-terminal to the catalytic domain, as the binding site for src family protein tyrosine kinases (66). EGF receptors (EGFR) bind PLCy with high affinity through the SH2 domain. It has been shown that the activated EGF R associates with She protein in vivo, most likely through its SH2 domain (67). The Shc protein serves as an adaptor protein which mediates the function of ras through other proteins such as Grb2 and Son of sevenless (803) (62). GAP and its associated proteins p62 and p190 become tyrosine and serine phosphorylated following stimulation with EGF (68), although a physical complex between GAP and the EGFR has not been identified. The in vitro studies have detected complexes between isolated GAP SH2 domains and the C-tail domain of the EGFR (69). Activation of PI3 kinase by the EGFR differs depending on the cell type (55). In pheochromocytoma cell line PC12, PI3 kinase is activated by the EGFR with kinetics comparable to those of other growth factor receptors (70). The EGFR has also been shown to function through G proteins (71). In human epiderrnoid carcinoma A431 cells, growth is inhibited by EGF, and EGF treatment of these cells induces tyrosine l6 phosphorylation and activation of signal transducers and activators of transcription (STATS) (72), a family of proteins that once activated, translocate from the cytoplasm to the nucleus and bind to DNA (73,74). The activation of STATS is mediated by another family of proteins called Janus kinases (JAK) that have been shown to associate with the EGF R (75). Many immediate early responsive genes including the cyclin-dependent kinase inhibitor p21 are regulated by JAK-STATs (75). The observation that a single receptor can interact with such a diverse range of signaling molecules demonstrates the diversity of the pathways that ligand binding can induce. For example, the association and regulation of PDGF receptor or EGF receptor with PI3 kinase, rasGAP, src family, PLC, JAK-STATs, etc. leads to the involvement of many pathways or components after ligand binding. The PI3 kinase has been shown to activate p70 ribosomal s6 kinase (p70rsk) (21). RasGAP is a regulator of the ras protein which interacts with raf, and leads to MAPK activation which activates pp90rsk or directly regulates c- jun/fos transcription (76), or the activation of ras can activate jun kinase (JNK) by some unknown intermediators (76). The activation of the src/yes/fyn family leads to myc gene transcription (77). The activation of PLCy hydrolyzes phophotidylinositol bisphosphate (PIP2) to generate diacylglycerol (DAG) and inositol (1, 4, 5)-trisphosphate (1P3). DAG in turn activates PKC and 1P3 induces intracellular Ca2+ release (21). Many of these pathways crosstalk with each other through certain proteins. For example, the activation of PKC and the elevation of intracellular Ca2+ lead to MAPK activation (50,53). The activation of P13 kinase has been shown to activate MAPK through the ras pathway (78). The activation of src has been shown to activate the ras/MAPK pathway (78), and JAK-STATs are connected with MAPK in lymphocytes (79). Pathwag involvinLthe G protein-coupled receptors: The activated G protein- coupled receptors interact with various types of G proteins (44), and G proteins, in turn, interact with their specific downstream effectors (44,78,80-82), mainly through, PLC, cyclic adenosine 3', 5'-phosphate (CAMP), ras/raf, and ion channels, depending on the type of G proteins involved. One important feature of the G protein-coupled receptors is that the same receptor may couple to different G proteins in different cell types (80). For example, bradykinin- activated PLC is fully inhibited by pertussis toxin in monkey renal cells (MDCK)(80,83,84), is partially inhibited by pertussis toxin in mouse NIH3T3 fibroblasts (80,85,86), and is not inhibited by pertussis toxin in human epidermoid carcinoma (A431) cells (80,87,88). The types of G proteins involved could reflect the availability of specific G protein in the cell type studied, or the G protein pools that can be accessed by the activated receptor in that cell type upon stimulation (80). Also, multiple pathways can be generated by varying the concentration of the ligand (89). For example, in pancreatic acinar tissue, Trimble et al. (90) observed that a low concentration of secretin stimulated adenylate cyclase, but a high concentration stimulated the activation of PLC to hydrolyze PIP2, leading to the generation of 1P3 and induces Ca2+ mobilization. Chabre et al. (91) suggested that these two pathways were mediated by different G proteins upon stimulation by different concentrations of the ligand, i. e., that G, directly activates adenylate cyclase and most likely a member of the Gq class regulates the stimulation of PLC. Alternatively, Bimbaumer (92) suggested that the stimulation of PLC could also be caused by beta gamma subunits of G, , which generate an additional signal but only at higher concentration of the ligands. 2.2 G proteins G proteins consist of three polypeptides, an a subunit, a B subunit, and a 7 subunit (44,93). The a subunit binds and hydrolyzes guanosine 5'-triphosphate (GTP) to guanosine 5'-diphosphate (GDP), and the B and 7 subunits form a dimer. When GDP is bound, the a subunit associates with the By subunits to form an inactive heterotrimer that binds to the receptor. Both a and By subunits can bind to the receptor (44). When a growth hormone or growth factor binds to a receptor, the receptor becomes activated and undergoes a conformational change. The GDP bound a subunit responds to this conformational change in the receptor and also changes its conformation, which decreases the affinity of the a subunit to GDP ; GDP then disassociates from this subunit, and is replaced with GTP. Once GTP is bound, the a subunit assumes its active conformation and dissociates from the receptor and the By subunit (44). All isoforms of 0. subunits are GTPases. However the GTP hydrolysis rate differs from one type of the a subunit to another (94,95). The free a and By subunits each activate their target effectors. There are over 20 different G protein (1 subunits in mammalian cells (44). The proteins are divided into four major classes (families) based on their amino acid sequence similarity: as (1,, (1,, (1,, and o.l2 (44) . Each of these classes has several members, for example, a, and a,lf are in the (its family, a“, (1,141,, 0., are in the o.i family, a, ,ctll ,uM, and aware in the (1,, family, and an, (1B are in the al2 family (44). A specific G protein activates its specific downstream effectors. Some types of G proteins or (1 subunits can be modified by cholera toxin or pertussis toxin (44). The a, family, which activates adenylate cyclase and regulates ion channels upon stimulation, can be modified by cholera toxin. The o.i family of 19 G proteins, which activates cyclic GTP, phosphodiesterase, or PLC (44,80), inhibits adenylate cyclase, or regulates K+ and Ca2+ channels, can be modified by pertussis toxin. (1,, and (1,, are insensitive to either of these two toxins. In general 01,, activates PLC, and r1l2 regulates NaVK+ exchange (81). In addition to the signals generated by Go subunits, GBy subunits can also generate signals (82,96-98). Ito et al. (99) showed that overexpression of GBy stimulated the phosphorylation of MAPK and enhanced the kinase activity of c-raf-l , and coexpression of dominant negative ras inhibited GBy-induced MAPK phosphorylation. This indicates that ras/raf/MAPK pathway can be activated by GBy. It has been shown that adenylate cyclase has a binding site for the GBy subunits (100). Hawes et al. (78) showed that GBy mediated MAPK activation through a pathway involving PI3 kinase, and that the activation of P13 kinase in this pathway was upstream of Sos/ras. They also showed that GBy was able to activate proteins of the src family, which in turn, activate She/Grb2 and Sos, and mediate the activation of ras/ raf/MAPK pathway. Krapivinsky et al. (101) showed that GB'y regulated the K+ channel by directly binding to it. 2.3 Phospholipase ‘C As described above, both the tyrosine kinase-linked receptors and G protein-coupled receptors can activate PLC. This enzyme is responsible for hydrolyzation of the inositol lipid, such as PIP2 , and generation of two important intracellular second messengers (43,102), DAG, and 1P3. DAG is responsible for activating the isoenzymes of PKC (103). On the endoplasmic reticulum, 1P3 binds with the 1P3 receptor which releases Ca2+ from intracellular stores. The release of intracellular Ca2+ is responsible for many cellular events. 20 The main isoforms of PLC are PLCB, PLCy, PLCS, and PLCs (80,104). Each one of these represents a family of closely related molecules. Each of the subtypes is designated by adding Arabic numerals after the Greek letters such as PLCyl and PLCyZ (104). PLCy has an approximate molecular weight of 150 kDa, and PLCB and PLC 8 have a molecular weight of 85 kDa. When assayed in vitro with pure substrate, PLCB, y, and 8 have been shown to hydrolyze both phosphotidylinositide (P1) and PIP2 (80). The hydrolysis of P1P2 occurs at micromolar concentration of Ca2+, and that of PI occurs at millimolar Ca” concentration (80). The PLCe forms have molecular weights of 85-88 kDa, and can only hydrolyze PIP2 but not PI when assayed in vitro (105-107), but no sequence data has yet been obtained on molecules of this family. The previously identified PLCa (80), which has molecular weights of 57-70 kDa and shares no homology with other isoforms of PLC, is now known to be a placental protein disulfide isomerase (108,109), and therefore is no longer classified as a PLC. Although the overall amino acid sequence similarity between different PLC isoforms is low, there is two regions of homology, region X, 150 amino acids, and region Y, 240 amino acids, which are shared by PLCB, PLCy, and PLCB (80,104). Among the three isoforms, there are about 60% identity in region X and 40% identity in region Y (80,110). These regions seem likely to constitute the catalytic domain. The various isoforms of PLC are regulated differently. PLC'y is regulated by tyrosine phosphorylation (80,111). It contains SH2 and SH3 domains. The SH2 domain targets the molecule to tyrosine phosphorylated sequences present in other proteins such as tyrosine kinase-linked receptors for PDGF, EGF, etc.. When activated, both PDGF and EGF receptors have been shown to be associated with PLCy (45,46) by their SH2 domain and to phosphorylate PLCyl on its tyrosine and serine residues (112-115). However, some other 21 receptors with tyrosine kinase activity, such as the insulin receptor and the CSF—l receptor, do not phosphorylate PLCyl (46,116). PLCB is regulated by G protein-coupled receptors (80). Evidence suggests that the G proteins involved in regulating PLCB include two distinct types: pertussis toxin sensitive and pertussis toxin insensitive (104). The (1 subunits of the Gq class of G proteins have been demonstrated to be specific activators of the PLCB isoforrn (117). The specificity of the interaction between different Ga subunits and PLC has been assessed by introducing cDNAs corresponding to various Ga subunits into the monkey kidney cell line Cos-7 and measuring the amount of inositol phosphates formed after stimulation with ALF; (104,118). Transfection of cells with the Gaq or G01ll cDNA results in a marked increase in inositol phosphate formation. Cotransfection of Gerq (or Gritll ) cDNA and PLC-B1 cDNA causes even higher levels of inositol phosphate formation. The relative ability of members of the Gq subfamily to activate PLCB isozymes has also been detennined. Membranes from Cos-7 cells transfected with cDNAs corresponding to Gaq , GetH , Getl4 , or Go“, have been reconstituted with purified PLCB] or PLCB2 in the presence of GTPyS (104). All four members of the Gaq subfamily were found to stimulate PLCB] , with Gaq or Gall being most effective. Grit“5 was found to activate PLCB2 most effectively, while Gaq , Ga” , or Gritl4 showed much less stimulation. Therefore, there appears to be a specificity in the interaction of different members for the Gq subfamily with different PLCB effectors. This specificity may be important in generating tissue- or receptor- specific responses in vivo. Other PLC regulators include Ca2+ and PKC. It has been shown that Ca2+ affects PLC activation either directly or by modulating the receptor mediated responses (80). An 22 increase of intracellular Ca2+ appears to stimulate PLC activity (80). However, there is very little information about which isoforms are Ca2+-regulated or whether there is a specific isoform of PLC that is only sensitive to Ca“. PLCyl can be negatively regulated by PKC (80,119). When Jurkat cells were treated with PMA, PLCy was shown to be phosphorylated at Ser-1248. This phosphorylation did not allow the tyrosine phosphorylation of this protein, and therefore, negatively regulated its activity (119). 2.4 Protein kinase C PKC has been shown to be activated by various growth factors and hormones (53,120). Multiple isoforms have been isolated, and different cell lines or tissues have been found to express different isoforms (121). Furthermore, different stimuli or second messengers may determine which PKC isoforms become activated (121). The PKC isoforms are divided into three groups (121-123): conventional or classic PKC(cPKC) which includes a, B1, B11, and 7, new PKC or novel PKC (nPKC) which includes 6, e, n, and 0, and atypical PKCs(aPKC) which includes C and A. It has been shown that PKCa, BI, B11, 5, a, and l; are ubiquitously distributed, e. g., in brain, lung, spleen, thymus and skin (121). PKC'y has been found to be exclusively expressed in the central nervous system. And PKCn is strongly expressed in skin and lung. PKCO is predominantly expressed in skeletal muscle. PKCa, 6, a, and C, have been shown to be expressed in both mouse NIH3T3 fibroblasts and rat6 fibroblasts (121,124,125). PKCs have a protein kinase domain which is conserved in all PKC isoforms, and a regulatory domain which shows different structures in different groups of PKCs (126). In the regulatory domain, all PKCs have one (aPKC) or two (cPKC, nPKC) cysteine-rich 23 sequences called the C1 region. This region is the putative membrane binding region (122), through which DAG activates PKC (121). The cPKC has a C2 region which contains many acidic amino acids and is thought to participate in Ca2+ binding (121,122) even though there is no sequence motif representing a known Ca2+ binding site (such as an E-F hand). The nPKC and aPKC lack the C2 region. The activation of nPKCs is dependent on DAG but not Ca2+. Phosphatidylcholine (PtdCho)/PLC or Phospholipase D hydrolysis of PtdCho leading to the generation of DAG only will activate this group of PKCs (123). The activation of aPKCs is not dependent on either Ca2+ or DAG. They can be regulated by the Ptdlns(3,4,5)P3 pathway or the ceramide pathway (123). There is sufficient and convincing evidence (121) showing that the activation of cPKC requires: the generation of DAG and 1P3 from plasma membrane-associated P1P2 by the activation of PLC, the release of Ca2+ from the intracellular Ca2+ store by 1P3 stimulation, the binding of Ca2+ to the C2 region of PKC, and subsequent translocation of PKC to the plasma membrane where it is activated by DAG through its C1 region. The activation of cPKC also requires phosphatidylserine as a cofactor which is constitutively present in the membrane (121). Phorbol ester mimicks the action of DAG to activate PKC (121). Activation of cPKC by phorbol ester does not require Ca2+, however the presence of Ca2+ will lower the concentration of phorbol ester required for the full activation of PKC (121). Phorbol ester treatment leads to the activation of PKC in various cells including Chinese hamster ovary (CHO) cell line (103), mouse NIH 3T3 cell line (121), Swiss 3T3 cell line (53), and human fibroblasts (127,128). In these cell lines, prolonged treatment with phorbol ester downregulates the activation of many PKC isoforms (53,121). The same PKC isoform will have a different PMA sensitivity in different cell lines. For example, PKCo. in most cell lines 24 is PMA sensitive, but in pituitary gland cells (GH4C1), this isoform is not downregulated upon treatment with PMA (121,129). PKCB and 5 are PMA sensitive in most cell lines, whereas the sensitivity of PKCe and PKCC to PMA is quite controversial (121). Different PKC isoforms have been overexpressed in different cell lines (121). The overexpression of PKC isoforms is correlated with an enhanced expression of the nuclear transcriptional factors such as c-jun, c-fos, and c-myc, and leads to the abnormal growth of the cells (121). In rat 6 fibroblasts (130), the overexpression of PKCBI causes the cells to grow in a disorganized manner and to be able to form colonies in soft agar. Rat 6 cells overexpressing PKCBI are more susceptible to transformation by tranfecting an activated H-ras oncogene compared to non-PKC expressing cells (131). Activation of different isoforms of PKC can phosphorylate downstream effectors in vivo. In CHO cells, overexpression of PKCa or PKC8 leads to MAPK activation (103). It has also been shown that the activation of PKC mediates c-raf kinase activation through direct phosphorylation (132,133). Studies with fibroblasts overexpressing PKC have suggested that PKC functions in an agonist-stirnulating pathway leading to the activation of phospholipase D or phospholipase A2 (122). The intracellular 85 kDa phospholipase A2 contains consensus phosphorylation sites for PKC and MAPK (122,134). However, activation of PKC alone does not appear to be sufficient. Intermediary proteins such as phospholipase A2-activating protein (135,136) may take part in activating this enzyme. 2.5 Raf Growth factors, such as PDGF, activate downstream kinases, such as PKC, raf, and MAPK, through phosphorylation (120,130,133). The viral-raf (v-raf) oncogene was 25 originally identified as the transforming gene of a murine sarcoma virus strain 3611 (137). The protooncogenes c-raf (138) and A-raf were identified (139) by screening a mouse spleen cDNA library with v-raf DNA as a probe. There are three active raf genes have been found in human, mouse, and chicken, A-raf-l , B-raf, and c-raf-l (140). They are dispersed over different chromosomes and have been mapped to sites that are frequently altered in human tumors (141,142). The raf genes encode cytosolic phosphoproteins of similar size: c-raf-l, 74 kDa; A-raf-l , 68 kDa; B-raf, 73 kDa; and these raf proteins function as serine/threonine— specific protein kinases (140,142). All raf proteins share three highly conserved regions (CR1, CR2, CR3) embedded in variable sequences (143). The catalytic domain of the raf kinases resides in CR3, whereas CR1 and CR2 fimction as the regulatory part of the raf protein molecules (143-145). 'Ihese raf proteins are differentially expressed in tissues (142): c-raf-l is ubiquitously expressed, A-raf-l is predominantly expressed in urogenital tissues (139), and B-raf is most abundant in cerebrum and testis. These kinases are inactive in quiescent cells, and can be activated by growth factors. B-raf has been shown to be activated in mammalian pheochromocytoma PC12 cells after stimulation by nerve growth factor (146,147). C-raf-l kinase is positively-regulated by activated growth factor receptor kinases in a variety of fibroblastic (140,148), epithelial (149), and lymphoid cells (150,151). Raf kinases are located in the cytosol of unstimulated cells. Growth factors induce ras activation which causes binding to the raf protein (140). Thus, binding by ras translocates raf to the plasma membrane which is necessary, but not sufficient, for the activation of raf kinase activity (152). Activation of raf kinase activity involves phosphorylation of serine/threonine residues on raf, and in the case of c-raf-l, of additional tyrosine residues by non-receptor tyrosine kinases (152,153). In the case of B-raf, 26 but not c-raf-l, phosphorylation and activation can be observed in vitro using crude preparations of the kinase and farnesylated ras (154). The identity of a specific kinase activator of raf remains unknown (155). PKC has been suggested as a candidate, since studies show that PKC mediates c-raf-l kinase activation by direct phosphorylation ( 132,133). The activation of raf kinase leads to a characteristic shift in apparent molecular weight upon phosphorylation (13,150). A-raf shares several functional properties with c-raf- 1 including transforming activity, stimulation of the raf/MAPK pathway, and the ability of dominant negative versions to functionally block ras signaling (156). Once activated, raf kinases phosphorylate the MAPK kinase (MEK) on its serine residues resulting in its activation (155). Two serine residues of MEK at positions 218 and 222 become phosphorylated, and either is sufficient for activation (157). MEK is a dual specificity threonine/tyrosine kinase, which when catalytically active, phosphorylates and activates the MAPK. Unlike raf and MEK, MAPKs have numerous substrates, including p90 ribosomal S6 kinase, cytoplamic phospholipase A2, c-jun, c-foc, c-myc, etc. (155). Raf-1 may also directly regulate the activation of the cell cycle through control of the mammalian Cdc25 phosphatase (158). The Cdc25 family of phosphatases activates many cyclin-dependent kinases (cdks) by dephosphorylation of conserved threonine and tyrosine residues (155). Active cdks control progression through each phase of the cell cycle. Raf-1 has been shown to physically complex with human Cdc25 in vivo and in vitro (2296). Immunoprecipitates of active, baculovirus-produced raf-l protein phosphorylate and activate the phosphatase activity of recombinant CchS (158). However, it is not clear whether Cdc25 is phosphorylated by raf-l or another associated kinase such as MEK or MAPK (155). 27 2.6. Intracellular Ca2+ The cytosolic concentrations of free Ca2+ in normal resting mammalian cells are very low (0.1-1.0 uM) compared with the extracellular levels (1.0-2.0 mM) (29). Consequently, the entry of a small quantity of Ca2+ into cells by Ca2+ channels in plasma membranes or the release of Ca2+ from intracellular stores will result in a sharp increase in free cytosolic Ca2+ concentrations. The sudden rise in cytosolic Ca2+ is believed to play an important role in cell proliferation (159). This idea is supported by the observation that mitogens, such as serum (16,160), PDGF (159,161,162), bradykinin (163,164), and bombesin (165), induce an intracellular Ca2+ increase (within seconds and which lasts for minutes) as well as cell proliferation in normal cells such as fibroblasts and muscle cells. By blocking Ca2+ influx with lanthanum, Estacion and Mordan (159) showed that the selective inhibition of the sustained Ca2+ increase completely inhibited the progression of mouse fibroblasts into S phase upon PDGF exposure. Direct evidence showing that intracellular Ca2+ is linked to cell growth has been obtained in studies by Gill and colleagues (166,167). Their studies showed that the intracellular Ca2+ pool depletion by thapsigargin arrested hamster smooth muscle cells in culture in a Go-like quiescent state and blocked cell growth (see below for detail). Ca2+ channels There are two types of Ca2+ channels in plasma membrane: voltage- sensitive and voltage-insensitive (47). In general, in excitable cells, such as neuronal and muscle cells, predominantly exhibit voltage sensitive calcium channels that open after membrane depolarization (168,169). There are three types of voltage sensitive calcium channels: L, T, and N type. N type Ca2+ channels are shown to be present exclusively in neuronal cells. L and T type Ca2+ channels are shown to be present in other types of cells, 28 including mouse and human fibroblasts as demonstrated by the patch-clamp technique which measures the conductance pattern of the cells (159,161,170). In nonexcitable cells, such as fibroblasts and epithelial cells, Ca2+ not only passes through the L and T types of Ca2+ channels (159,161,170), but also a family of poorly characterized, voltage insensitive calcium channels (169,171), which have been classified into three general groups: 1) receptor-operated calcium channels, 2) second messenger-operated calcium channels, and 3) depletion-operated calcium channels. Regulation of the first group has been shown to be independent of intracellular Ca2+ and some other second messengers (172). For example, muscarinic receptors can regulate these Ca2+ channels (169), and the structural elements in the receptor plays a role in its regulation (172). Stimulation of the second group of channels can be directly regulated by intracellular calcium, 1P3, and possibly 1P4. The third group of channels open after 1P3-mediated depletion of intracellular stores and provide a source of calcium for refilling the stores. Regulation of calcium influx through this group of channels is triggered by the depletion of intracellular calcium and may involve a calcium influx factor released from the endoplasmic reticulum that stimulates the opening of low conductance calcium channels (168,169) It has been shown that Ca2+ entry through the different channels regulates gene expression differently (173). Bading ct al. (173) demonstrated that N-Methyl-D-aspartate (N MDA) receptors and L-type Ca2+ channels, the two major sites of Ca2+ entry into hippocampal neuronal cells, transmit signals to the nucleus and regulate gene transcription through two distinct Ca2+ signaling pathways. These investigators found that the L-type Ca2+ channels triggered Ca2+ entry and then initiated c-fos gene expression through the Ca2+ response element and/or the CAMP response element on the c-fos promoter region between 29 base pairs +1 and -222. Gene expression of c-fos was dependent on Ca2+-calmodulin- dcpcndent protein kinase (CaMK) (173). On the other hand, NMDA receptor Ca2+ channels induced the expression of c-fos through the serum response element of the c-fos promoter region located on or near base pair -300. In this case, gene expression did not depend on CaMK. These investigators also found that a specific L-type channel blocker or a NMDA receptor antagonist completely abolished c-fos expression (173). Intracellular Ca“ pools There are multiple intracellular Ca2+ pools present in endoplasmic reticulum of various cell systems (174): the 1P3 sensitive pool and the 1P3 insensitive pool, both of which are sensitive to thapsigargin, and a third pool is unresponsive to 1P3 and insensitive to thapsigargin. The thapsigargin sensitive pools (pools 1 and 2) contain 75% of the total Ca2+ , and the third pool contains about 25% (174). Baurngarten et al. (175) showed that human foreskin fibroblasts also has three different intracellular Ca2+ pools: an 1P3 sensitive, thapsigargin sensitive pool; an 1P3 insensitive, thapsigargin sensitive pool; and an ionomycin sensitive pool which is not sensitive to either 1P3 or thapsigargin. Studies by Gill and colleagues (166,167) have shown that thapsigargin depletes the intracellular Ca2+ stores (pools) by specific inhibition of the endoplasmic reticulum (ER) Ca2+ -ATPasc, and in turn, will deplete the source of intracellular Ca2+ for these pools. They found that the depletion of those Ca2+ pools resulted in the arrest of cells in G0 phase (166,174), and thapsigargin suppressed the expression of the endoplasmic reticulum Ca2+ pump protein (176). 2.7 Calmodulin 30 The function of Ca2+ is mediated by its binding proteins. There are two different kinds of Ca2+ binding proteins, trigger proteins and buffer proteins (168). Buffer proteins, such as calsequestn'n, simply bind Ca2+ as its concentration increases within cells to buffer or lower free Ca2+ levels since too high a free Ca2+ concentration is toxic to cells (168). Trigger proteins, on another hand, change their conformation upon binding Ca2+ and modulate effector molecules such as enzyme and ion channels. The trigger proteins include calmodulin, calcineurin, phospholipase A2, PLC,, etc. (168). Calmodulin acts as the primary mediator of Ca2+-dependent signaling in eukaryotic nonmuscle and smooth muscle cells by serving as a high affinity intracellular receptor (177,178). It has been shown that calmodulin is required for progression at specific points of the cell cycle in mammalian cells (179,180) and that it affects cell proliferation in variety of cell types (1,181). For example, the selective pharmacological inhibitors of calmodulin kinase, KN -62 and KN-93, have been shown to prevent quiescent WS-l human fibroblasts from reaching S phase upon serum stimulation (181,182). Takuwa et al. (16) showed that the potent calmodulin antagonists calmidazolium and N-(6-aminohexyl)-5-chloro-l-naphthalenesulfonamide HCl (W-7) but not its inactive analogue N-(4-aminobutyl)-2-naphthalenesulfonamide (W-12), strongly inhibit serum- induced DNA synthesis in normal human fibroblasts (16) and human vascular cndothelial cells (183). In neuronal cells, it has been shown (173) that the increase in intracellular Ca2+ caused by Ca2+ entry modulates the activation of calmodulin kinases (CaMK), such as CaMK2 or CaMK 4, regulating the expression of their downstream genes. Calmodulin is highly conserved throughout the eukaryotic kingdoms (184). It has a molecular weight of approximately 17 kDa and a high percentage of acidic amino acids. A distinctive property of calmodulin and of other related calcium-binding proteins is that they undergo a mobility 31 shift when run on a SDS-PAGE in the presence of Ca2+ (184). Calmodulin has been shown to be encoded by several genes in vertebrates. Three calmodulin cDNAs (CaMl, CaM2, and CaM3) have been cloned from rat and human, and multiple genes have been identified in many other species (178,184,185). These cDNAs all encode identical calmodulin proteins. Although all three mRN A transcripts are expressed in all mammalian tissues and cultured cells, the expression level of each calmodulin gene has been shown to vary according to the cell type and is modified by the concentration of extracellular growth factors, such as nerve growth factor (186). In mouse C127 cells, it has been shown (178) that all three calmodulin genes are expressed, but that only the CaM2 mRNA levels show significant changes as cells progress through the cell cycle. These results indicate that the calmodulin genes may be differentially regulated during the cell cycle. In mammalian cells, the 148 amino acid calmodulin protein has a dumbbell-shape structure with two Ca2+ binding sites in each half of the molecule (178). Except for the budding yeast calmodulin which only binds three Ca2+ ions, calmodulins from all other species have four highly conserved “BF-hand” Ca2+-binding sites, like those first described in the crystal structure of parvaburnin (178,187). When intracellular Ca2+ increases, it binds to calmodulin, and induces a conformational change in the molecule which exposes its hydmphobic patches that are involved in interaction with and activation of its target enzymes (177,184,188,189). Without the conformational alterations, calmodulin does not bind to most of its various targets (184). More than 20 enzymes can be regulated by calmodulin (178,188,190,191). These enzymes include cyclic 3', 5'-nucleotide phosphodiesterase (178,192,193), adenylate cyclase (194,195), (Ca2+- Mg2+)ATPase (196,197), the cardiac microsomal calcium transporter (197), 1P3 kinase (178,198,199), calmodulin-dependent 32 protein kinases, such as myosin light chain kinase (178,200) and the multifunctional calmodulin-dependent protein kinase (201,202) as well as calmodulin-dependent protein phosphatase (calcineurin) (203-205). Ca2+ has been shown to be absolutely required for all enzyme activating functions of vertebrate calmodulin (178). However, the Ca2+ requirement can be altered in vitro by different concentrations of calmodulin. Increasing the calmodulin concentration can decrease the amount of Ca2+ required to activate calmodulin-dependent enzymes. Also, increasing the Ca2+ concentration can decrease the amount of calmodulin required for the activation of calmodulin-dependent enzymes (178). These results indicate that in vitro Ca2+ and calmodulin cooperatively regulate the function of the target protein. Transformed cells typically have elevated calmodulin levels and exhibit the ability to grow in Ca2+-deficient medium which does not allow the growth of their nontransforrned counterparts (178). Because it has not yet been possible to replace all three active endogenous calmodulin genes with a single inducible calmodulin gene, the relationship between the calmodulin concentration and the Ca2+ requirement for cell growth has remained unclear. It has been shown that the growth of A. nidulans cells is dependent on the concentration of calmodulin and Ca2+ as in mammalian cells (206). Using A. nidulans as a model system, it has been shown (178) that increasing the calmodulin concentration allows these cells to grow at a very low extracellular Ca2+ concentration (IO-fold lower than the noninduced or normal state). These data indicate that a cooperative regulation exists inside cells between Ca2+ and calmodulin, and this may provide a possible explanation for why cells that are transformed and have elevated calmodulin levels can proliferate in a Ca2+-deficient medium (178). 33 2.8 MAPK As described above, protein growth factors such as EGF, PDGF, insulin, thrombin, nerve growth factor (N GF ), and bombesin, can initiate a phosphorylation cascade in various cell types through the activation of either tyrosine kinase-linked or G protein-coupled receptors, and one of the key events in this cascade is the activation of MAPK by phosphorylation (49-54,207,208). Several closely related MAPK isoforms, usually referred to by their molecular masses, i.e., p40, p42, p44, and p54 (209), have been identified in various cell types. In human fibroblasts, it has been shown that EGF (50), tumor necrosis factor (52,210,211) and interleukin (211) can cause activation of p42 and/or p44 MAPKs. MAPK kinase activates MAPK by phosphorylation on its tyrosine and serine/threonine residues (212), which results in a mobility shift of MAPK proteins when analyzed by SDS- PAGE. The activated MAPK can, in turn, phosphorylate microtubule-associated protein (MAP-2) (213), myelin basic protein (MBP) (214), ribosomal S6 kinase (215), the EGF receptor (216), phospholipase A2 (217), and the nuclear transcriptional factors c-myc (218,219), and c-fos and c-jun (216,220). Pouysségur and his colleagues (221,222) showed that in quiescent Chinese hamster fibroblast cell line CCL39, MAPK was activated by thrombin or basic fibroblast growth factor and that the activation of MAPK (as detected by an immunocomplex kinase assay using MBP as substrate) exhibited a biphasic pattern i.e., a rapid increase in activity followed by a rapid decrease and then a sustained level of lower activity. They proposed that this sustained activity was required for the re-entry of these cells from G0 into the cell cycle (222), i.e., for triggering the proliferative response, because 34 when they treated CCL39 cells with a thrombin analog instead of thrombin or with carbachol, the second phase of MAPK activation was not observed and no DNA synthesis took place. Other investigators have demonstrated that the sustained activation of MAPK is associated with proliferation in fibroblasts (223-225). In addition, Pages et al. (226) and Frost et al. (227), using MAPK antisense RNA and/or MAPK kinase-deficient mutants to suppress MAPK activation, showed that MAPK activation was required for activation of transcriptional factors (227) and for rodent fibroblast proliferation (226). In other cell systems, continuous MAPK activation could lead to the opposite cellular behavior. In pheochromocytoma cell line PC12, cellular responses are determined by the duration of MAPK activation. In these cells, it has been shown that sustained MAPK activation is associated with differentiation, whereas transient MAPK activation is related to proliferation (228). Continuous treatment of PC12 cells with FGF or NGF leads to outgrowth of neurites and eventual cessation of cell division, whereas treatment with EGF leads to a proliferative signal (229). Many signal transduction events have been found to be shared between differentiation and proliferation in PC12 cells (230). However, notable quantitative differences were found between differentiation and proliferative signals. In PC12 cells, NGF stimulation results in a persistent elevation of rasGTP, whereas EGF produces only a short lived rise in rasGTP (231). MAPK activation is sustained for several hours following NGF stimulation, but it is short lived after EGF stimulation (229,232). In PC12 cells, it has been found that sustained MAPK activation is always associated with translocation of MAPK to the nucleus (229,232-234), whereas transient activation does not lead to nuclear translocation. Transient activation will, therefore, have very different 35 consequences for gene expression compared with sustained activation (2314). In this way, it is believed that cells can use transient and sustained activation of MAPK to determine a variety of responses (228). C. Ca2+ signaling in mammalian cells 1. Intracellular components involved in Ca2+ signaling and cell differentiation/hormonal secretion As described in “Section A”, extracellular Ca2+ can serve as the first messenger to regulate cell differentiation, hormonal secretion, and cell proliferation in different cell systems (29-3 5,23 5). The numerous intracellular components or pathways involved in these processes have been found using different approaches (29-31,34,235) In keratinocytes in culture, the addition of Ca2+ to the medium rapidly increases the intracellular phosphoinositide turnover, intracellular Ca2+ levels (30,34), and tyrosine phosphorylation of proteins (30,31). The rasGAP has been shown to be specifically affected during Ca2 +-induced differentiation (31). Upon Ca2+ treatment, GAP associates with tyrosine-phosphorylated proteins and translocates to the membrane (31). Also, a p62 protein, which is associated with GAP, has been found to be heavily phosphorylated on tyrosine both in membrane and cytosolic fractions (31). This protein is only phosphorylated by Ca2+ stimulation, and not via growth factor stimulation, such as EGF or TGF-B (30,31). Ca2 1 does not induce the tyrosine phosphorylation of PLCy and PI3 kinase, which are activated upon EGF stimulation in this cell type (31). Filvaroff ct al. (30) suggested that Ca2+-induced p62 phosphorylation, and possibly differentiation, was mediated by a Ca2+ receptor. Denning et 36 al. (235) showed that the induction of keratinocyte differentiation markers by Ca2+ was mediated by specific protein isozymes. In the cultured murine cells they studied (235), they detected PKC isoforms by immunoblotting (PKCu, 8, e, C). They observed that PKCa, 8, and a were translocated from the soluble fraction to particulate fraction during Ca2+ -induced differentiation, suggesting that the induction of keratinocyte differentiation by Ca2+ results in the activation of specific PKC isoforms (235). In parathyroid cells, the increase of extracellular Ca“ concentration decreases PTH release (29). Studies (236,237) showed that stepwise increases in the extracellular Ca2+ concentration produced corresponding increases in intracellular Ca2+ concentration in these cells. There was a close inverse relationship between the high extracellular Ca2+-induced intracellular Ca2+ increase and the accompanying decreases in PTH release (236,237), although it is not clear yet whether the intracellular Ca2+ change is the cause of the decrease in PTH release. However, other studies (29,23 8) showed that increasing the intracellular Ca2+ directly by treating cells with Ca” ionophores, such as A23 1 87, was associated with inhibition of PTH secretion. Extracellular Ca2+ also stimulates the accumulation of inositol monophosphate, inositol bisphosphate, 1P3, and IP-4 (239). It is believed that the increases of 1P3 mediates, at least partially, the intracellular Ca2+ elevation evoked by extracellular Ca2+ (240). Shoback et al. (239) showed that Ca2+ ionophore, ionomycin, which raises the intracellular Ca2+ directly to a level equivalent to that observed at high extracellular Ca2+ concentration, had little or no effect on the accumulation of inositol phosphates (29). And the addition of extracellular Ca2+ to parathyroid cells increases 1P3 level, which has been shown to be through the activation of PLC, and is mediated by a G protein-coupled receptor (29,241). In addition, extracellular Ca2+ inhibits the receptor-mediated cAMP increase. The decrease of cAMP level upon Ca2+ treatment is 37 mediated by enhanced adenylate cyclase activity (2,29,242). In osteoclasts, extracellular Ca2+ stimulates Ca2 1 release from Ca2 1 entry channels in the plasma membrane which can be blocked by lanthanum (37) and from intracellular stores. A Ca2+ receptor is suggested to be involved in the increase of intracellular Ca“ in this cell type (37). They showed that in osteoclast cells, extracellular Ca2 1 induced intracellular Ca2+ elevation, which leads to cytoskeletal changes affecting cell adhesion and decreasing the bone resorptive activity (3 7). 2. Intracellular components involved in Ca2+ stimulated growth pathway A few groups have studied the mechanism of Ca2 +-induced cell growth. Studies by Rubin and Samui (9) and Bowen Pope and Rubin (10) showed that in mouse fibroblasts the Ca2 + and phosphate precipitate can serve as the primary signal to promote cell growth. In their studies, they observed that Ca2+-stimulated DNA synthesis was correlated with the concentration of inorganic orthophosphate or pyrophosphate in the medium, as well as the turbidity of the medium. They proposed that precipitates of Ca2+-phosphate complexes acted in a manner similar to macromolecular stimulants (such as hormones) and interacted with the plasma membrane. Most recently, Epstein et al. (13) and Epstein (14) showed that calcium chloride mimicked PDGF in its ability to stimulate DNA synthesis and to induce c- fos gene expression in mouse 3T3 fibroblasts. They showed that both PDGF-induced and Ca2+-induced c-fos expression is dependent on PKC, but unlike PDGF, Ca2+ does not induce the phosphorylation of the PDGF receptor and the raf protein, indicating that Ca2+ and PDGF have distinct pathways resulting in similar gene transcription. 38 3. Ca2+ sensing receptor and its signaling In 1993, Brown et al. (241) cloned a Ca2+-sensing receptor from bovine parathyroid, and subsequently, the human Ca2+-sensing receptor was cloned (243). This receptor possesses an extracellular domain of about 613 amino acids, a membrane domain which contains seven membrane-spanning segments, and an intracellular domain of about 222 amino acids) (241,244). The seven membrane-spanning domain is the characteristic of the superfamily of G-protein coupled receptors as described above in section B2.2. This protein is about 120 kDa (241,244) and it has been shown to be expressed both in tissues known to be involved in Ca2+ homeostasis (e.g., parathyroid, C-cells, and kidney) and in other cells that do not directly regulate systemic Ca2+ (e.g., brain, neuronal, and smooth muscle cells) (2). The Ca2+-sensing receptor has been shown to transmit signals from outside the cell to induce cellular responses, such as to increase cytosolic Ca“ and [PB levels when expressed in xenopus oocyte (241). The rise of intracellular Ca2+ is sustained by the Ca2+ influx through Ca2+ channels in the plasma membrane (2). And also, the expression of this receptor cDNA in xenopus oocyte has been shown to activate PLC (241,244,245) which will hydrolyze PIP2 to generate 1P3 and induce intracellular Ca2+ elevation. The activation is dependent on a G protein (244,246) which is sensitive to pertussis toxin in xenopus oocyte (241). However, the elevation of 1P3 and intracellular Ca2+ in parathyroid cells is not dependent on pertussis toxin sensitive G protein (2,29), indicating that this receptor couples to a different subfamily of G protein in these different cell types (241) It has been shown that human diseases, such as familial hypocalciuric hypercalcemia and neonatal severe hyperparathyroidism, result from mutations in this Ca2+-sensing receptor (247-249). Various mutants of this receptor have been generated to study the function of this 39 protein (249-252). The results from these studies confirm that the mutation of this receptor is responsible for various human diseases. In parathyroid cells in culture, it has been shown that the reduction or loss of Ca2+ responsiveness is related to a marked reduction in expression of extracellular Ca2+-sensing receptor (253,254). Recently, knockout mouse lacking Ca2+-sensing receptor (255) has been generated and it has the characteristics similar to those of humans with a mutation in this receptor. The heterozygous mice (CaR+/-) are analogous to humans with familial hypocalciuric hypercalcemia and have modest elevations of serum calcium, magnesium and parathyroid hormone levels, as well as hypocalciuria. The CaR-l- mice are analogous to humans with neonatal severe hyperparathyroidism. They have markedly elevated serum calcium and parathyroid hormone levels, parathyroid hyperplasia, bone abnormalities, as well as retarded growth (255). This suggests that Ca2+ receptor mutations cause these human disorders by reducing the number of firnctional receptor molecules on the cell surface (255). 4. Other possible Ca’I-sensing receptors In parathyroid cells, extracellular Ca2+ not only induces the intracellular 1P3 and Ca2+ increase, but also inhibits receptor-mediated increases in CAMP (2,29). For example, dopamine-stimulated CAMP is inhibited by extracellular Ca” (29). This inhibition can be abolished by pertussis toxin in parathyroid cells (29), indicating that a pertussis toxin- sensitive G protein is involved in this process (2,29). Such an inhibiting effect on CAMP is also observed in the thick ascending limb cell from mouse kidney (256). The decrease of CAMP level upon Ca2+ treatment is due to the enhanced adenylate cyclase activity (2,29,242), which is activated via a pertussis toxin-sensitive G protein. This G protein apparently differs 40 from that which mediates the Ca2+-induced elevation of intracellular 1P3 and intracellular Ca2+. The receptor coupled to this pertussis toxin-sensitive G protein in parathyroid cells is either the same receptor Cloned by Brown et al.(2) which couples to two different G protein subfamilies simultaneously (2), or could be another receptor which senses extracellular Ca2+ and mediates this action (29). By using monoclonal antibodies, Juhlin et al. (257-259) identified an ~500 kDa membrane protein on proximal tubular cells, cytotrophoblast cells of human placenta. Exposure of these cells to extracellular Ca2+ (2) induces cellular responses, such as regulation of l-hydroxylation of 25-hydroxyvitamin D in the proximal tubule and Changes in circulating PTH levels. The fact that the Ca2+-sensing receptor cloned by Brown et al. from bovine parathyroid has not been found to be expressed in these cells (2) suggests that the 500 kDa protein is another type of Ca2+ sensing receptor. Recently, Lundgren et al. (260) isolated a 2.8 kb cDNA which contains an open reading fiame encoding 236 amino acids, and which represents part of the gene form the above mentioned 500 kDa protein. Using this cDNA as probe, a 15 kb transcript was identified in human kidney, placenta, and parathyroid (260). Sequence analysis of the 2.8 kb cDNA showed that this protein is a member of the low- density lipoprotein receptor superfamily and that it is Closely related to Heyman nephritogenic protein, a kidney tubule glycoprotein, with Ca2+ binding activity (260,261). The Heyman nephritogenic protein has been Cloned and sequenced recently (262). It is the human variant of principal kidney autoantigen causing Heyman membranous glomcrulonephritis in rats (262). The deduced 4655 amino acid residues give a calculated molecular weight of 519.6 kDa for the mature protein. The primary structure reveals that it contains an huge extracellular region (about 4400 amino acids), a single transmembrane- 4l spanning domain of 23 amino acids, and an intracellular C-terminal domain of 209 amino acids (262). This protein contains three types of cysteine-rich repeats Characteristic of the low density lipoprotein receptor (LDLR) superfamily (262), and its intracellular domain contains several SH3 recognition motifs, one SH2 recognition motif for the P85 regulatory subunit of PI3 kinase, and additional recognition sites for PKC, casein kinase II, and CAMP/CGMP dependent protein kinase (262). It is not yet known whether this protein, Heyman nephritogenic protein, or the 500 kDa protein identified by Juhlin et al. (257-259), function as Ca2+ sensors. Subsequent functional studies should resolve this question. In addition to specific Ca2+ sensing receptors, other known growth factor receptors could also mediate the function of extracellular Ca” and generate intracellular signals. For example, it has been shown that the F GF receptor has a Ca2+ binding motif in its acid box, which binds Ca2+ as shown by a Ca2+ blotting technique, and the Ca2+ binding is not observed in a mutagenized form of the receptor that lacks the acidic box region (263). The mechanism(s) by which Ca2+ directly modulates the activities of various cell types (2), such as keratinocytes (15), cultured mammary cells (264), and intestinal goblet cells (265) still needs to be identified. 42 LIST OF REFERENCES 1. Bootrnan, MD. and Ben'idge, M.J. The elemental principles of calcium signaling. Cell, 83: 675-678, 1995. 2. Hebert, SC. and Brown, E.M. The extracellular calcium receptor. Curr. Opin. Cell Biol., 7: 484-492, 1995. 3. Balk, S.D., Polimeni, P.1., Hoon, B.S., LeStourgeon, D.N., and Mitchell, R.S. Proliferation of Rous sarcoma virus-infected, but not of normal, Chicken fibroblasts in a medium of reduced calcium and magnesium concentration. . Proc. Natl. Acad. Sci. U. S. A., 76: 3913-3916, 1979. 4. Balk, S.D., Whitfield, J.F., Youdale, T., and Braun, A.C. Roles of calcium, scrum, plasma, and folic acid in the control of proliferation of normal and Rous sarcoma virus-infected Chicken fibroblasts. Proc. Natl. Acad. Sci. U. S. A., 70: 675-679, 1973. 5. Balk, S.D. Calcium as a regulator of the proliferation of normal, but not of transformed, Chicken fibroblasts in a plasma-containing medium. Proc. Natl. Acad. Sci. U. S. A., 68: 271-275, 1971. 6. Dulbecco, R. and Elkington, J. Induction of growth in resting fibroblastic cell cultures by Ca++. Proc. Natl. Acad. Sci. U. S. A., 72: 1584-1588, 1975. 7. Hazelton, B., Mitchell, B., and Tupper, J. Calcium, magnesium, and growth control in the WI-38 human fibroblast cell. J. Cell Biol., 83 : 487-498, 1979. 8. Barnes, D.W. and Colowick, S.P. Stimulation of sugar uptake and thymidine incorporation in mouse 3T3 cells by calcium phosphate and other extracellular particles. Proc. Natl. Acad. Sci. U. S. A., 74: 5593-5597, 1977. 9. Rubin, H. and Sanui, H. Complexes of inorganic pyrophosphate, orthophosphate, and calcium as stimulants of 3T3 cell multiplication. Proc. Natl. Acad. Sci. U. S. A., 74: 5026-5030, 1977. 10. Bowen Pope, DP. and Rubin, H. Growth stimulatory precipitates of Ca2+ and pyrophosphate. J. Cell Physiol., 11 7: 51-61, 1983. 11. Mitchell, P.G., Pledger, W.J., and Cheung, H.S. Molecular mechanism of basic calcium phosphate crystal-induced mitogenesis. Role of protein kinase C. J. Biol. Chem, 264: 14071-14077, 1989. 12. Cheung, H.S., Mitchell, P.G., and Pledger, W.J. Induction of expression of c-fos and c-myc protooncogenes by basic calcium phosphate crystal: effect of beta-interferon. . Cancer 43 Res., 49: 134-138, 1989. 13. Epstein, R.J., Druker, B.J., Irrninger, J.C., Jones, S.D., Roberts, T.M., and Stiles, C.D. Extracellular calcium mimics the actions of platelet-derived growth factor on mouse fibroblasts. Cell Growth Differ., 3: 157-164, 1992. 14. Epstein, R. Calcium-inducible transmodulation of receptor tyrosine kinase activity. Cell Signal., 7: 377-388, 1995. 15. Hennings, H., Michael, D., Cheng, C., Steinert, P., Holbrook, K., and Yuspa, S.H. Calcium regulation of growth and differentiation of mouse epidermal cells in culture. Cell, 19: 245-254, 1980. 16. Takuwa, N., Zhou, W., Kumada, M., and Takuwa, Y. Ca(2+)-dependcnt stimulation of retinoblastoma gene product phosphorylation and p34cdc2 kinase activation in serum-stimulated human fibroblasts. J. Biol. Chem, 268: 138-145, 1993. 17. Praeger, F .C. and Cristofalo, V.J. The growth of WI-38 cells in a serum-free, growth factor-free, medium with elevated calcium concentrations. In. Vitro. Cell. Dev. Biol., 22: 355-359, 1986. 18. Praeger, F .C. and Cristofalo, V.J. Modulation of WI-38 cell proliferation by elevated levels of CaC12. J. Cell. Physiol., 129: 27-35, 1986. 19. Pledger, W.J., Stiles, C.D., Antoniadcs, H.N., and Scher, C.D. Induction of DNA synthesis in BALB/c 3T3 cells by serum components: reevaluation of the commitment process. Proc. Natl. Acad. Sci. U. S. A., 74: 4481-4485, 1977. 20. Stiles, C.D., Capone, G.T., Scher, C.D., Antoniadcs, H.N., Van Wyk, J.J., and Pledger, W.J. Dual control of cell growth by somatomedins and platelet—derived growth factor. Proc. Natl. Acad. Sci. U. S. A., 76: 1279-1283, 1979. 21. Benidge, M.J. Calcium signalling and cell proliferation. Bioessays, 1 7: 491-500, 1995. 22. van der Burg, B., Isbrucker, L., van Selm Miltenburg, A.J., de Laat, S.W., and van Zoelen, E.J. Role of estrogen-induced insulin-like growth factors in the proliferation of human breast cancer cells. Cancer Res., 50: 7770-7774, 1990. 23. Werner, S., Roth, W.K., Bates, B., Goldfarb, M., and Hofschneider, P.H. Fibroblast growth factor 5 proto-oncogene is expressed in normal human fibroblasts and induced by serum growth factors. Oncogene, 6: 2137-2144, 1991. 24. Sugimoto, T., Kanatani, M., Kano, J., Kobayashi, T., Yamaguchi, T., F ukase, M., and Chihara, K. IGF-I mediates the stimulatory effect of high calcium concentration on osteoblastic cell proliferation. Am. J. Physiol., 266: E709-E716, 1994. 44 25. Morgan, T.L., Yang, D.J., Fry, D.G., Hurlin, P.J., Kohler, S.K., Maher, V.M., and McCormick, J .J . Characteristics of an infinite life span diploid human fibroblast cell strain and a near-diploid strain arising from a clone of cells expressing a transfected v-myc oncogene. Exp. Cell Res., 197: 125-136, 1991. 26. Bettger, W.J., Boyce, S.T., Walthall, B.J., and Ham, R.G. Rapid Clonal growth and serial passage of human diploid fibroblasts in a lipid-enriched synthetic medium supplemented with epidermal growth factor, insulin, and dexarnethasonc. Proc. Natl. Acad. Sci. U. S. A., 78: 5588-5592, 1981. 27. Ryan, P.A., Maher, V.M., and McCormick, J.J. Modification of MCDB 110 medium to support prolonged growth and consistent high Cloning efficiency of diploid human fibroblasts. Exp. Cell Res., 172: 318-328, 1987. 28. Wilson, D.M., Yang, D.J., Dillberger, J.E., Dietrich, S.E., Maher, V.M., and McCormick, J .J . Malignant transformation of human fibroblasts by a transfected N-ras oncogene. Cancer. Res., 50: 5587-5593, 1990. 29. Brown, E.M. Extracellular Ca2+ sensing, regulation of parathyroid cell function, and role of Ca2+ and other ions as extracellular (first) messengers. Physiol. Rev., 71: 371-411, 1991. 30. Filvaroff, E., Calautti, E., Reiss, M., and Dotto, G.P. Functional evidence for an extracellular calcium receptor mechanism triggering tyrosine kinase activation associated with mouse keratinocyte differentiation. . J. Biol. Chem, 269: 2173 5-21 740, 1994. 31. Filvaroff, E., Calautti, E., McCormick, F., and Dotto, G.P. Specific Changes of Ras GTPase-activating protein (GAP) and a GAP-associated p62 protein during calcium-induced keratinocyte differentiation. Mol. Cell Biol., 12: 5319-5328, 1992. 32. DiGiovanni, J ., Gill, R.D., Nettikumara, A.N., Colby, AB, and Reincrs, J.J.J. Effect of extracellular calcium concentration on the metabolism of polycyclic aromatic hydrocarbons by cultured mouse keratinocytes. Cancer Res., 49: 5567-5574, 1989. 33. Yuspa, S.H., Hennings, H., Tucker, R.W., Jaken, S., Kilkcnny, A.E., and Roop, D.R. Signal transduction for proliferation and differentiation in keratinocytes. Ann. N. Y. Acad. Sci., 548: 191-196, 1988. 34. Tang, W., Ziboh, V.A., Isseroff, R, and Martinez, D. Turnover of inositol phospholipids in cultured murinc keratinocytes: possible involvement of inositol triphosphate in cellular differentiation. J. Invest. Dermatol., 90: 37-43, 1988. 35. Nygren, P., Larsson, R., Johansson, H., Gylfe, E., Rastad, J., and Akerstrom, G. Inhibition of cell growth retains differentiated firnction of bovine parathyroid cells in 45 monolayer culture. Bone. Miner., 4: 123-132, 1988. 36. Sakaguchi, K. Acidic fibroblast growth factor autocrine system as a mediator of calcium-regulated parathyroid cell growth. J. Biol. Chem, 26 7: 24554-24562, 1992. 37. Miyauchi, A., Hruska, K.A., Greenfield, B.M., Duncan, R., Alvarez, J., Barattolo, R., Colucci, S., Zambonin Zallone, A., Teitelbaum, S.L., and Teti, A. Osteoclast cytosolic calcium, regulated by voltage-gated calcium Channels and extracellular calcium, controls podosome assembly and bone resorption. J. Cell Biol., 111: 2543-2552, 1990. 38. Choy, VI. and Watkins, W.B. Effect of ionic environment on the release of human placental lactogen in vitro. J. Endocrinol., 69: 349-358, 1976. 39. Handwerger, S., Conn, P.M., Barrett, J., Barry, 8., and Golander, A. Htunan placental lactogen release in vitro: paradoxical effects of calcium. Am. J. Physiol., 240: E550-E555, 1981. 40. Massague, J. and Pandiella, A. Membrane-anchored growth factors. Annu. Rev. Biochem, 62: 515-541, 1993. 41. Heldin, C.H. Dimerization of cell surface receptors in signal transduction. Cell., 80: 213-223, 1995. 42. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., and Watson, J .D. Cell Signaling. In: M. Robertson (ed.), Molecular Biology of the Cell, third edition, pp. 721-785, New York: Garland publishing,lnc.. 1994. 43. Berridge, M.J. Inositol trisphosphate and calcium signaling. Ann. N. Y. Acad. Sci., 766: 3 1-43, 1995. 44. Neer, E.J. Heterotrimeric G proteins: organizers of transmembrane signals. Cell, 80: 249-257, 1995. 45. Kumjian, D.A., Wahl, M.I., Rhee, S.G., and Daniel, T.O. Platelet-derived growth factor (PDGF) binding promotes physical association of PDGF receptor with phospholipase C. Proc. Natl. Acad. Sci. U. S. A., 86: 8232-8236, 1989. 46. Nishibe, S., Wahl, M.I., Wedegaertner, P.B., Kim, J .W., Rhee, S.G., Carpenter, G., and Kim, J .J .K. Selectivity of phospholipase C phosphorylation by the epidermal growth factor receptor, the insulin receptor, and their cytoplasmic domains [published erratum appears in Proc Natl Acad Sci U S A 1990 Apr;87(8):3253]. Proc. Natl. Acad. Sci. U. S. A., 87: 424-428, 1990. 47. Trofunova, M., Sprenkle, A.B., Green, M., Sturgill, T.W., Goebl, M.G., and Harrington, M.A. Developmental and tissue-specific expression of mouse pelle-like protein kinase. 1. 46 Biol. Chem, 271: 17609-17612, 1996. 48. Cockcroft, S. Relationship between arachidonate release and exocytosis in permeabilized human neutrophils stimulated with forrnylmethionyl-leucyl-phenylalanine (fMetLeuPhe), guanosine 5'-[gamma-thio]triphosphate (GTP[S]) and Ca2+. Biochem. J., 275: 127-131, 1991. 49. Burgering, B.M., de Vries Smits, A.M., Medema, R.H., van Weeren, P.C., Tertoolen, LG, and Bos, J .L. Epidermal growth factor induces phosphorylation of extracellular signal-regulated kinase 2 via multiple pathways. Mol. Cell Biol., 13: 7248-7256, 1993. 50. Chao, T.S., Byron, K.L., Lee, K.M., Villereal, M., and Rosner, M.R. Activation of MAP kinases by calcium-dependent and calcium-independent pathways. Stimulation by thapsigargin and epidermal growth factor. J. Biol. Chem, 26 7: 19876-19883, 1992. 51. Nakielny, S., Cohen, P., Wu, J., and Sturgill, T. MAP kinase activator from insulin-stimulated skeletal muscle is a protein threonine/tyrosine kinase. EMBO. J., 11: 2123-2129, 1992. 52. Victor, 1., Schwenger, R, Li, W., Schlessinger, J., and Vilcek, J. Tumor necrosis factor-induced activation and increased tyrosine phosphorylation of mitogen-activated protein (MAP) kinase in human fibroblasts. J. Biol. Chem, 268: 18994-18999, 1993. 53. Pang, L., Decker, S.J., and Saltiel, A.R. Bombesin and epidermal growth factor stimulate the mitogen-activated protein kinase through different pathways in Swiss 3T3 cells. . Biochem. J., 289: 283-287, 1993. 54. Lenormand, P., Sardet, C., Pages, G., L'Allemain, G., Brunet, A., and Pouyssegur, J. Growth factors induce nuclear translocation of MAP kinases (p42mapk and p44mapk) but not of their activator MAP kinase kinase (p45mapkk) in fibroblasts. J. Cell Biol., 122: 1079-1088, 1993. 55. F antl, W.J., Johnson, DE, and Williams, L.T. Signalling by receptor tyrosine kinases. Annu. Rev. Biochem, 62: 453-481, 1993. 56. Moran, M.F., Koch, C.A., Anderson, D., Ellis, C., England, L., Martin, GS, and Pawson, T. Src homology region 2 domains direct protein-protein interactions in signal transduction. Proc. Natl. Acad. Sci. U. S. A., 87: 8622-8626, 1990. 57. Anderson, D., Koch, C.A., Grey, L., Ellis, C., Moran, M.F., and Pawson, T. Binding of SH2 domains of phospholipase C gamma 1, GAP, and SrC to activated grth factor receptors. Science, 250: 979-982, 1990. 58. Koch, C.A., Anderson, D., Moran, M.F., Ellis, C., and Pawson, T. SH2 and SH3 domains: elements that control interactions of cytoplasmic signaling proteins . Science, 252: 47 668-674, 1991. 59. Fantl, W.J., Escobedo, J.A., Martin, G.A., Turck, C.W., del Rosario, M., McCormick, F., and Williams, L.T. Distinct phosphotyrosines on a growth factor receptor bind to specific molecules that mediate different signaling pathways. . Cell., 69: 413-423, 1992. 60. Reedijk, M., Liu, X., van der Geer, P., Letwin, K., Waterficld, M.D., Hunter, T., and Pawson, T. Tyr721 regulates specific binding of the CSF-l receptor kinase insert to P1 3'-kinase SH2 domains: a model for SH2-mediated receptor-target interactions. EMBO. J ., 11: 1365-1372, 1992. 61. Backer, J .M., Myers, M.G.J., Shoelson, S.E., Chin, D.J., Sun, X.J., Miralpeix, M., Hu, P., Margolis, B., Skolnik, E.Y., Schlessinger, J ., and et al , Phosphatidylinositol 3'-kinase is activated by association with IRS-l during insulin stimulation. EMBO. J., 11: 3469-3479, 1992. 62. Fry, M.J., Panayotou, G., Booker, G.W., and Waterfield, MD. New insights into protein-tyrosine kinase receptor signaling complexes. Protein Sci., 2: 1785-1797, 1993. 63. Kazlauskas, A., Kashishian, A., Cooper, J .A., and Valius, M. GTPase-activating protein and phosphatidylinositol 3-kinase bind to distinct regions of the platelet-derived growth factor receptor beta subunit. Mol. Cell. Biol., 12: 2534-2544, 1992. 64. Kashishian, A. and Cooper, J.A. Phosphorylation sites at the C-terrninus of the platelet-derived growth factor receptor bind phospholipase C gamma 1 . Mol. Biol. Cell., 4: 49-57, 1993. 65. F eng, G.S., Hui, CC, and Pawson, T. SH2-containing phosphotyrosine phosphatase as a target of protein-tyrosine kinases . Science, 259: 1607-1611, 1993. 66. Mori, S., Ronnstrand, L., Yokote, K., Engstrom, A., Courtneidge, S.A., Claesson Welsh, L., and Heldin, C.H. Identification of two juxtarnembrane autophosphorylation sites in the PDGF beta-receptor; involvement in the interaction with Src farnily tyrosine kinases . EMBO. J., 12: 2257-2264, 1993. 67. Pelicci, G., Lanfiancone, L., Grignani, F., McGlade, J ., Cavallo, F ., Fomi, G., Nicoletti, 1., Pawson, T., and Pelicci, P.G. A novel transforming protein (SHC) with an SH2 domain is implicated in mitogenic signal transduction . Cell., 70: 93-104, 1992. 68. Ellis, C., Moran, M., McCormick, F ., and Pawson, T. Phosphorylation of GAP and GAP-associated proteins by transforming and mitogenic tyrosine kinases. . Nature, 343: 377-381, 1990. 69. Margolis, 3., Li, N., Koch, A., Mohammadi, M., Hurwitz, D.R., Zilberstein, A., Ullrich, A., Pawson, T., and Schlessinger, J. The tyrosine phosphorylated carboxyterminus of the 48 EGF receptor is a binding site for GAP and PLC-gamma . EMBO. J ., 9: 4375-4380, 1990. 70. Carter, AN. and Downes, C.P. Phosphatidylinositol 3-kinase is activated by nerve growth factor and epidermal growth factor in PC12 cells [published erratum appears in J Biol Chem 1992 Nov 15;267(32):23434]. J. Biol. Chem, 267: 14563-14567, 1992. 71. Poppleton, H., Sun, H., Fulgham, D., Bertics, P., and Patel, T.B. Activation of Gsalpha by the epidermal growth factor receptor involves phosphorylation. J. Biol. Chem, 2 71: 6947-6951, 1996. 72. Chin, Y.E., Kitagawa, M., Su, W.C., You, Z.H., Iwamoto, Y., and Fu, X.Y. Cell growth arrest and induction of cyclin-dependent kinase inhibitor p21 WAF l/CIPl mediated by STAT]. Science, 272: 719-722, 1996. 73. Sadowski, H.B., Shuai, K., Darnell, J.E.J., and Gilman, M.Z. A common nuclear signal transduction pathway activated by growth factor and cytokinc receptors [see comments] . Science, 261: 1739-1744, 1993. 74. Ihle, J .N. STATS; signal transducers and activators of transcription. Cell., 84: 331-334, 1996. 75. Ihle, J.N., Witthuhn, B.A., Quelle, F.W., Yarnarnoto, K., Thierfelder, W.E., Kreider, B., and Silvennoinen, O. Signaling by the cytokine receptor superfamily: JAKs and STATS . Trends Biochem. Sci., 19: 222-227, 1994. 76. Davis, R.J. MAPKs: new JNK expands the group . Trends Biochem. Sci., 19: 470-473, 1994. 77. Benito, M. and Lorenzo, M. Platelet derived growth factor/tyrosine kinase receptor mediated proliferation . Grth Regul., 3: 172-179, 1993. 78. Hawes, B.E., Luttrell, L.M., van Biesen, T., and Leflcowitz, R.J. Phosphatidylinositol 3-kinase is an early intermediate in the G beta garnma-mediated mitogen-activated protein kinase signaling pathway . J. Biol. Chem, 271: 12133-12136, 1996. 79. David, M., Petricoin, E., Benjamin, C., Pine, R., Weber, M.J., and Larner, A.C. Requirement for MAP kinase (ERK2) activity in interferon alpha-and interferon beta-stimulated gene expression through STAT proteins [see comments] . Science, 269: 1721-1723, 1995. 80. Cockcroft, S. and Thomas, G.M. Inositol-lipid-specific phospholipase C isoenzymes and their differential regulation by receptors . Biochem. J., 288: 1-14, 1992. 81. Voyno Yasenetskaya, T., Conklin, B.R., Gilbert, R.L., Hooley, R., Boume, HR, and Barber, D.L. G alpha 13 stimulates Na-H exchange. J. Biol. Chem, 269: 4721-4724, 1994. 49 82. Kim, D., Lewis, D.L., Graziadei, L., Neer, E.J., Bar Sagi, D., and Clapham, D.E. G-protein beta garnma-subunits activate the cardiac muscarinic K+-channel via phospholipase A2. Nature, 33 7: 557-560, 1989. 83. Portilla, D., Morrissey, J., and Morrison, A.R. Bradykinin-activated membrane-associated phospholipase C in Madin-Darby canine kidney cells. J. Clin. Invest, 81: 1896-1902, 1988. 84. Slivka, SR. and lnsel, P.A. Phorbol ester and neomycin dissociate bradykinin receptor-mediated arachidonic acid release and polyphosphoinositide hydrolysis in Madin-Darby canine kidney cells. Evidence that bradykinin mediates noninterdependent activation of phospholipases A2 and C. J. Biol. Chem, 263: 14640-14647, 1988. 85. Hoshijirna, M., Ueda, T., Harnamori, Y., Ohmori, T., and Takai, Y. Different sensitivity to phorbol esters and pertussis toxin of bombesin- and platelet-derived growth factor-induced, phospholipase C-mediated hydrolysis of phosphoinositides in NIH/3T3 cells. Biochem. Biophys. Res. Commun., 152: 285-293, 1988. 86. Fu, T., Okano, Y., and Nozawa, Y. Bradykinin—induced generation of inositol 1,4,5-trisphosphate in fibroblasts and neuroblastoma cells: effect of pertussis toxin, extracellular calcium, and down-regulation of protein kinase C. Biochem. Biophys. Res. Commun., 157: 1429-1435, 1988. 87. Tilly, B.C., van Paridon, P.A., Verlaan, I., de Laat, S.W., and Moolenaar, W.H. Epidermal-growth-factor-induced formation of inositol phosphates in human A431 cells. Differences from the effect of bradykinin. Biochem. J., 252: 857-863, 1988. 88. Tilly, B.C., van Paridon, P.A., Verlaan, 1., Wirtz, K.W., de Laat, S.W., and Moolenaar, W.H. Inositol phosphate metabolism in bradykinin-stirnulated human A431 carcinoma cells. Relationship to calcium signalling. Biochem. J., 244: 129-135, 1987. 89. Bygrave, F .L. and Roberts, H.R. Regulation of cellular calcium through signaling cross-talk involves an intricate interplay between the actions of receptors, G-proteins, and second messengers. FASEB. J., 9: 1297-1303, 1995. 90. Trimble, E.R., Bruzzone, R., Biden, T.J., Meehan, C.J., Andreu, D., and Merrifield, R.B. Secretin stimulates cyclic AMP and inositol trisphosphate production in rat pancreatic acinar tissue by two fully independent mechanisms. Proc. Natl. Acad. Sci. U. S. A., 84 : 3146-3150, 1987. 91. Chabre, 0., Conklin, B.R., Lin, H.Y., Lodish, H.F., Wilson, E., Ives, H.B., Catanzariti, L., Hemmings, B.A., and Boume, H.R. A recombinant calcitonin receptor independently stimulates 3',5'-cyclic adenosine monophosphate and Ca2+/inositol phosphate signaling pathways [see comments]. . Mol. Endocrinol., 6: 551-556, 1992. 50 92. Bimbaumer, L. Receptor-to-effector signaling through G proteins: roles for beta gamma dimers as well as alpha subunits. Cell, 71: 1069-1072, 1992. 93. Neer, E.J. and Smith, T.F. G protein heterodimers: new structures propel new questions. Cell, 84: 175-178, 1996. 94. Carty, DJ. and Iyengar, R. A 43 kDa form of the GTP-binding protein GB in human erythrocytes . FEBS Lett., 262: 101-103, 1990. 95. Linder, M.E., Ewald, D.A., Miller, R.J., and Gilman, A.G. Purification and characterization of Go alpha and three types of Gi alpha after expression in Escherichia coli. J. Biol. Chem, 265: 8243-8251, 1990. 96. Clapham, DE. and Neer, E.J. New roles for G-protein beta gamma-dimers in transmembrane signalling . Nature, 365: 403-406, 1993. 97. Wickman, K.D., Iniguez Lluhl, J.A., Davenport, P.A., Taussig, R., Krapivinsky, G.B., Linder, M.E., Gilman, A.G., and Clapham, D.E. Recombinant G-protein beta gamma-subunits activate the muscarinic-gated atrial potassium Channel. Nature, 368: 255-257, 1994. 98. Krapivinsky, G., Krapivinsky, L., Wickman, K., and Clapham, D.E. G beta gamma binds directly to the G protein-gated K+ Channel, IKACh. J. Biol. Chem, 270: 29059-29062, 1995. 99. Ito, A., Satoh, T., Kaziro, Y., and Itoh, H. G protein beta gamma subtmit activates Ras, Raf, and MAP kinase in HEK 293 cells. FEBS Lett., 368: 183-187, 1995. 100. Simon, M.I., Strathmann, M.F., and Gautam, N. Diversity of G proteins in signal transduction. Science, 252: 802-808, 1991. 101. Maltese, WA. and Robishaw, J.D. Isoprenylation of C-tenninal cysteine in a G-protein gamma subunit. J. Biol. Chem, 265: 18071-18074, 1990. 102. Berridge, M.J. Inositol trisphosphate and calcium signalling. Nature, 361: 315-325, 1993. 103. Yamaguchi, K., Ogita, K., Nakamura, S., and Nishizuka, Y. The protein kinase C isoforms leading to MAP-kinase activation in CHO cells. Biochem. Biophys. Res. Commun., 210: 639-647, 1995. 104. Rhee, S.G. and Choi, K.D. Regulation of inositol phospholipid-specific phospholipase C isozymes. J. Biol. Chem, 267: 12393-12396, 1992. 51 105. Rebecchi, M.J. and Rosen, O.M. Purification of a phosphoinositide-specific phospholipase C from bovine brain. . J. Biol. Chem, 262: 12526-12532, 1987. 106. Fukui, T., Lutz, R.J., and Lowenstein, J.M. Purification of a phospholipase C from rat liver cytosol that acts on phosphatidylinositol 4,5-bisphosphate and phosphatidylinositol 4-phosphate. J. Biol. Chem, 263: 17730-17737, 1988. 107. Homma, Y., Imaki, J ., Nakanishi, 0., and Takenawa, T. Isolation and characterization of two different forms of inositol phospholipid-specific phospholipase C from rat brain. J. Biol. Chem, 263 : 6592-6598, 1988. 108. Chamock Jones, D.S., Day, K., and Smith, S.K. Cloning, expression and genomic organization of human placental protein disulfidc isomerase (previously identified as phospholipase C alpha). Int. J. Biochem. Cell Biol., 28: 81-89, 1996. 109. Srivastava, S.P., Fuchs, J.A., and Holtzrnan, J.L. The reported cDNA sequence for phospholipase C alpha encodes protein disulfide isomerase, isozyme Q-2 and not phospholipase-C. Biochem. Biophys. Res. Commun., 193: 971-978, 1993. 110. Rhee, S.G., Kim, H., Suh, P.G., and Choi, W.C. Multiple forms of phosphoinositide-specific phospholipase C and different modes of activation. Biochem. Soc. Trans, 19: 337-341, 1991. 111. Rhee, S.G. Inositol phospholipids-specific phospholipase C: interaction of the gamma 1 isoform with tyrosine kinase. Trends. Biochem. Sci., 16: 297-301, 1991. 112. Meisenhelder, J., Suh, P.G., Rhee, S.G., and Hunter, T. Phospholipase C-gamma is a substrate for the PDGF and EGF receptor protein-tyrosine kinases in vivo and in vitro. Cell., 57: 1109-1122, 1989. 113. Margolis, B., Rhee, S.G., Felder, S., Mervic, M., Lyall, R., Levitzki, A., Ullrich, A., Zilberstein, A., and Schlessinger, J. EGF induces tyrosine phosphorylation of phospholipase C-II: a potential mechanism for EGF receptor signaling. Cell., 5 7: 1101-1107, 1989. 114. Nishibe, S., Wahl, M.I., Rhee, S.G., and Carpenter, G. Tyrosine phosphorylation of phospholipase C-II in vitro by the epidermal growth factor receptor. J. Biol. Chem, 264 : 10335-10338, 1989. 115. Wahl, M.I., Nishibe, S., Suh, P.G., Rhee, S.G., and Carpenter, G. Epidermal growth factor stimulates tyrosine phosphorylation of phospholipase C-II independently of receptor internalization and extracellular calcium. Proc. Natl. Acad. Sci. U. S. A., 86: 1568-1572, 1989. 116. Downing, J.R., Margolis, B.L., Zilberstein, A., Ashmun, R.A., Ullrich, A., Sherr, C.J., and Schlessinger, J. Phospholipase C-gamma, a substrate for PDGF receptor kinase, is not 52 phosphorylated on tyrosine during the mitogenic response to CSF-l. EMBO. J., 8: 3345-3350, 1989. 117. Park, D., Jhon, D.Y., Lee, C.W., Lee, K.H., and Rhee, S.G. Activation of phospholipase C isozymes by G protein beta gamma subunits. J. Biol. Chem, 268: 4573-4576, 1993. 118. Lee, C.H., Park, D., Wu, D., Rhee, S.G., and Simon, M.I. Members of the Gq alpha subunit gene farnily activate phospholipase C beta isozymes. J. Biol. Chem, 26 7: 16044-16047, 1992. 119. Park, D.J., Min, H.K., and Rhee, S.G. Inhibition of CD3-linked phospholipase C by phorbol ester and by CAMP is associated with decreased phosphotyrosine and increased phosphoserine contents of PLC-gamma 1. J. Biol. Chem, 26 7: 1496-1501, 1992. 120. Finkenzeller, G., Marme, D., Weich, HA, and Hug, H. Platelet-derived growth factor-induced transcription of the vascular endothelial growth factor gene is mediated by protein kinase C. Cancer Res., 52: 4821-4823, 1992. 121. Hug, H. and Sarre, T.F. Protein kinase C isoenzymes: divergence in signal transduction? Biochem. J., 291 : 329-343, 1993. 122. Nishizuka, Y. Intracellular signaling by hydrolysis of phospholipids and activation of protein kinase C. Science, 258: 607-614, 1992. 123. Divecha, N. and Irvine, R.F. Phospholipid signaling. Cell, 80: 269-278, 1995. 124. Li, W., Michieli, P., Alimandi, M., Lorenzi, M.V., Wu, Y., Wang, L.H., Heidaran, M.A., and Pierce, J.H. Expression of an ATP binding mutant of PKC-delta inhibits Sis-induced transformation of NIH3T3 cells . Oncogene, 13: 731-737, 1996. 125. Hirai, S., Izurni, Y., Higa, K., Kaibuchi, K., Mizuno, K., Osada, S., Suzuki, K., and Ohno, S. Ras-dependcnt signal transduction is indispensable but not sufficient for the activation of APl/Jun by PKC delta. EMBO J., 13: 2331-2340, 1994. 126. Sasaki, Y., Asaoka, Y., and Nishizuka, Y. Potentiation of diacylglycerol-induced activation of protein kinase C by lysophospholipids. Subspecics difference . F EBS Lett., 320: 47-51, 1993. 127. Bi, N. and Mamrack, M.D. PMA inhibits the growth of human fibroblasts after the induction of immediate-early genes . Exp. Cell Res., 212: 105-112, 1994. 128. Hendey, B. and Mamrack, M.D. WS-l human fibroblasts contain distinct calcium and protein kinase C-mediated pathways for activation of Na+/H+ exchange: contrasting effects of thrombin and PMA. J. Cell Physiol., 146: 290-297, 1991. 53 129. Akita, Y., Ohno, S., Yajima, Y., and Suzuki, K. Possible role of Ca2(+)-independent protein kinase C isozyme, nPKC epsilon, in thyrotropin—releasing hormone-stimulated signal transduction: differential down-regulation of nPKC epsilon in GH4C1 cells. Biochem. Biophys. Res. Commun., 172: 184-189, 1990. 130. Housey, G.M., Johnson, M.D., Hsiao, W.L., O'Brian, C.A., Murphy, J.P., Kirschmeier, P., and Weinstein, I.B. Overproduction of protein kinase C causes disordered growth control in rat fibroblasts. . Cell, 52: 343-354, 1988. 131. Hsiao, W.L., Housey, G.M., Johnson, M.D., and Weinstein, 1.3. Cells that overproduce protein kinase C are more susceptible to transformation by an activated H-ras oncogene. Mol. Cell Biol., 9: 2641—2647, 1989. 132. Kolch, W., Heidecker, G., Kochs, G., Hummel, R., Vahidi, H., Mischak, H., Finkenzeller, G., Manne, D., and Rapp, U.R. Protein kinase C alpha activates RAF-l by direct phosphorylation. Nature, 364: 249-252, 1993. 133. Sozeri, 0., Vollmer, K., Liyanage, M., Frith, D., Kour, G., Mark, GE, and Stabel, S. Activation of the c-Raf protein kinase by protein kinase C phosphorylation. Oncogene, 7: 2259-2262, 1992. 134. Clark, J.D., Lin, L.L., Kriz, R.W., Ramesha, C.S., Sultzrnan, L.A., Lin, A.Y., Milona, N., and Knopf, J.L. A novel arachidonic acid-selective cytosolic PLA2 contains a Ca(2+)-dependent translocation domain with homology to PKC and GAP. Cell, 65: 1043-1051,1991. 135. Bomalaski, J.S., Baker, D.G., Brophy, L., Resurreccion, N.V., Spilbcrg, 1., Muniain, M., and Clark, M.A. A phospholipase A2-activating protein (PLAP) stimulates human neutrophil aggregation and release of lysosomal enzymes, superoxide, and eicosanoids. J. Immunol., 142: 3957-3962, 1989. 136. Clark, M.A., Ozgur, L.E., Conway, T.M., Dispoto, J., Crooke, ST, and Bomalaski, J.S. Cloning of a phospholipase A2-activating protein. Proc. Natl. Acad. Sci. U. S. A., 88: 5418-5422, 1991. 137. Rapp, U.R., Goldsborough, M.D., Mark, G.E., Bonner, T.I., Groffen, J., Reynolds, F.H.J., and Stephenson, J.R. Structure and biological activity of v-raf, a unique oncogene transduced by a retrovirus. Proc. Natl. Acad. Sci. U. S. A., 80: 4218-4222, 1983. 138. Bonner, T.I., Oppermann, H., Seeburg, P., Kerby, S.B., Gunnell, M.A., Young, AC, and Rapp, U.R. The complete coding sequence of the human raf oncogene and the corresponding structure of the c-raf-l gene. Nucleic. Acids. Res., 14: 1009-1015, 1986. 139. Huleihel, M., Goldsborough, M., Cleveland, J., Gunnell, M., Bonner, T., and Rapp, U.R. Characterization of murine A-raf, a new oncogene related to the v-raf oncogene. Mol. 54 Cell. Biol., 6: 2655-2662, 1986. I40. Rapp, U.R. Role of Raf-l serine/threonine protein kinase in growth factor signal transduction. Oncogene, 6: 495-500, 1991. 141. Storm, S.M., Brennscheidt, U., Sithanandam, G., and Rapp, U.R. raf oncogenes in carcinogenesis. Crit. Rev. Oncog., 2: 1-8, 1990. 142. Storm, S.M., Cleveland, J .L., and Rapp, U.R. Expression of raf family proto-oncogenes in normal mouse tissues. Oncogene, 5: 345-351, 1990. 143. Avruch, J., Zhang, X.F., and Kyriakis, J.M. Raf meets Ras: completing the framework of a signal transduction pathway. Trends. Biochem. Sci., 19: 279-283, 1994. 144. Rapp, U.R., Heidecker, G., Huleihel, M., Cleveland, J.L., Choi, W.C., Pawson, T., Ihle, J.N., and Anderson, W.B. raf family serine/threonine protein kinases in mitogen signal transduction. Cold. Spring. Harb. Symp. Quant. Biol., 53 Pt 1: 173-184, 1988. 145. Heidecker, G., Huleihel, M., Cleveland, J.L., Kolch, W., Beck, T.W., Lloyd, P., Pawson, T., and Rapp, U.R. Mutational activation of C-raf-l and definition of the minimal transforming sequence. Mol. Cell. Biol., 10: 2503-2512, 1990. 146. Vaillancourt, R.R., Gardner, A.M., and Johnson, G.L. B-Raf-dependent regulation of the MEK-I/mitogen-activatcd protein kinase pathway in PC12 cells and regulation by cyclic AMP. . Mol. Cell. Biol., 14: 6522-6530, 1994. 147. Jaiswal, R.K., Moodic, S.A., Wolfrnan, A., and Landreth, GB. The mitogen-activated protein kinase cascade is activated by B-Raf in response to nerve growth factor through interaction with p2lras. Mol. Cell. Biol., 14: 6944-6953, 1994. 148. Morrison, D.K., Kaplan, D.R., Escobedo, J.A., Rapp, U.R., Roberts, T.M., and Williams, L.T. Direct activation of the serine/threonine kinase activity of Raf-1 through tyrosine phosphorylation by the PDGF beta-receptor. Cell., 58: 649-657, 1989. 149. Yin, T., Keller, S.R., Quelle, F.W., Witthuhn, B.A., Tsang, M.L., Lienhard, G.E., Ihle, J.N., and Yang, Y.C. Interleukin-9 induces tyrosine phosphorylation of insulin receptor substrate-l via JAK tyrosine kinases. J. Biol. Chem, 270: 20497-20502, 1995. 150. Siegel, J.N., Klausner, R.D., Rapp, U.R., and Samelson, L.E. T cell antigen receptor engagement stimulates C-raf phosphorylation and induces c-raf-associated kinase activity via a protein kinase C-dependent pathway. . J. Biol. Chem, 265 : 18472-18480, 1990. 151. Siegel, J.N., June, C.H., Yamada, H., Rapp, U.R., and Samelson, L.E. Rapid activation of C-Raf-l after stimulation of the T-ccll receptor or the muscarinic receptor type 1 in resting T cells. J. Immunol., 151: 4116-4127, 1993. 55 152. Marais, R., Light, Y., Paterson, H.F., and Marshall, C.J. Ras recruits Raf-l to the plasma membrane for activation by tyrosine phosphorylation. EMBO. J ., 14: 3136-3145, 1995. 153. Dent, P., Reardon, D.B., Morrison, D.K., and Sturgill, T.W. Regulation of Raf-1 and Raf-l mutants by Ras-dependent and Ras-indcpendent mechanisms in vitro [published erratum appears in Mol Cell Biol 1995 Sep;15(9):5203]. Mol. Cell Biol., 15: 4125-4135, 1995. 154. Yamamori, B., Kuroda, S., Shimizu, K., Fukui, K., Ohtsuka, T., and Takai, Y. Purification of a Ras-dependent mitogen-activated protein kinase kinase kinase from bovine brain cytosol and its identification as a complex of B-Raf and 14-3-3 proteins. J. Biol. Chem, 270: 11723-11726, 1995. 155. Marshall, M.S. Ras target proteins in eukaryotic cells. FASEB. J., 9: 1311-1318, 1995. 156. Lee, J.E., Beck, T.W., Wojnowski, L., and Rapp, U.R. Regulation of A-raf expression. Oncogene, 12: 1669-1677, 1996. 157. Daum, G., Eisenmann Tappe, 1., Fries, H.W., Troppmair, J., and Rapp, U.R. The ins and outs of Rafkinascs. . Trends. Biochem. Sci., 19: 474-480, 1994. 158. Galaktionov, K., Jessus, C., and Beach, D. Rafl interaction with Cdc25 phosphatase ties mitogenic signal transduction to cell cycle activation. Genes. Dev., 9: 1046-1058, 1995. 159. Estacion, M. and Mordan, L.J. Expression of voltage-gated calcium channels correlates with PDGF-stimulated calcium influx and depends upon cell density in C3H 10T1/2 mouse fibroblasts. Cell Calcium, 14: 161-171, 1993. ’ 160. Takuwa, N., Zhou, W., Kumada, M., and Takuwa, Y. Involvement of intact inositol-1,4,5-trisphosphate-sensitive Ca2+ stores in cell cycle progression at the Gl/S boundary in serum-stimulated human fibroblasts. F EBS Lett., 360: 173-176, 1995. 161. Wang, Z., Estacion, M., and Mordan, L.J. Ca2+ influx via T-type Channels modulates PDGF-induced replication of mouse fibroblasts. Am. J. Physiol., 265 : C1239-C1246, 1993. 162. Estacion, M. and Mordan, L.J. Competence induction by PDGF requires sustained calcium influx by a mechanism distinct from storage-dependent calcium influx. Cell Calcium, 14: 439-454, 1993. 163. Huang, H.M., Toral Barza, L., and Gibson, G.E. Interactions between inositol phosphates and cytosolic free calcium following bradykinin stimulation in cultured human skin fibroblasts. Biochim. Biophys. Acta., 1091: 409-416, 1991. 56 164. Byron, K.L., Babnigg, G., and Villereal, M.L. Bradykinin-induced Ca2+ entry, release, and refilling of intracellular Ca2+ stores. Relationships revealed by image analysis of individual human fibroblasts. . J. Biol. Chem, 26 7: 108-118, 1992. 165. Takuwa, N., Iwamoto, A., Kumada, M., Yamashita, K., and Takuwa, Y. Role of Ca2+ influx in bombesin-induced mitogenesis in Swiss 3T3 fibroblasts. J. Biol. Chem, 266: 1403-1409, 1991. 166. Ghosh, T.K., Bian, J.H., Short, A.D., Rybak, S.L., and Gill, D.L. Persistent intracellular calcium pool depletion by thapsigargin and its influence on cell growth. . J. Biol. Chem, 266: 24690-24697, 1991. 167. Short, A.D., Bian, J., Ghosh, T.K., Waldron, R.T., Rybak, S.L., and Gill, D.L. Intracellular Ca2+ pool content is linked to control of cell growth. Proc. Natl. Acad. Sci. U. S. A., 90: 4986-4990, 1993. 168. Clapham, D.E. Calcium signaling. Cell, 80: 259-268, 1995. 169. Felder, C.C. Muscarinic acetylcholine receptors: signal transduction through multiple effectors. FASEB. J., 9: 619-625, 1995. 170. Chen, C.F., Corbley, M.J., Roberts, T.M., and Hess, P. Voltage-sensitive calcium Channels in normal and transformed 3T3 fibroblasts. Science, 239: 1024-1026, 1988. 171. Felder, C.C., Singer Lahat, D., and Mathes, C. Voltage-independent calcium channels. Regulation by receptors and intracellular calcium stores. Biochem. Pharmacol, 48: 1997-2004, 1994. 172. Felder, C.C., Poulter, M.O., and Wess, J. Muscarinic receptor-operated Ca2+ influx in transfected fibroblast cells is independent of inositol phosphates and release of intracellular Ca2+. Proc. Natl. Acad. Sci. U. S. A., 89: 509-513, 1992. 173. Bading, H., Ginty, DD, and Greenberg, M.E. Regulation of gene expression in hippocampal neurons by distinct calcium signaling pathways. Science, 260: 181-186, 1993. 174. Bian, J.H., Ghosh, T.K., Wang, J.C., and Gill, D.L. Identification of intracellular calcium pools. Selective modification by thapsigargin. J. Biol. Chem, 266: 8801-8806, 1991. 175. Baumgarten, L.B., Lee, HQ, and Villereal, M.L. Multiple intracellular Ca2+ pools exist in human foreskin fibroblast cells: the effect of BK on release and filling of the non-cytosolic Ca2+ pools. Cell Calcium, 17: 41-52, 1995. 176. Waldron, R.T., Short, A.D., Meadows, J.J., Ghosh, T.K., and Gill, D.L. Endoplasmic reticulum calcium pump expression and control of cell growth. J. Biol. Chem, 269: 57 11927-11933, 1994. 177. Means, A.R., VanBerktun, M.F., Bagchi, 1., Lu, K.P., and Rasmussen, C.D. Regulatory functions of calmodulin. Pharrnacol. Ther., 50: 255-270, 1991. 178. Lu, K.P. and Means, A.R. Regulation of the cell cycle by calcium and calmodulin. Endocr. Rev., 14: 40-58, 1993. 179. Rasmussen, CD. and Means, A.R. Calmodulin is involved in regulation of cell proliferation. . EMBO. J., 6: 3961-3968, 1987. 180. Rasmussen, CD. and Means, A.R. Calmodulin is required for cell-cycle progression during G1 and mitosis. EMBO. J., 8: 73-82, 1989. 181. Takuwa, N., Zhou, W., and Takuwa, Y. Calcium, calmodulin and cell cycle progression. Cell Signal., 7: 93-104, 1995. 182. Minami, H., Inoue, S., and Hidaka, H. The effect of KN-62, Ca2+/calmodulin dependent protein kinase II inhibitor on cell cycle. Biochem. Biophys. Res. Commun., 199: 241-248, 1994. 183. Takuwa, N., Zhou, W., Kumada, M., and Takuwa, Y. Ca2+/calmodulin is involved in growth factor-induced retinoblastoma gene product phosphorylation in human vascular endothelial cells. FEBS Lett., 306: 173-175, 1992. 184. Hinrichsen, R.D. Calcium and calmodulin in the control of cellular behavior and motility. . Biochim. Biophys. Acta., 1155: 277-293, 1993. 185. Nojima, H. Structural organization of multiple rat calmodulin genes. J. Mol. Biol., 208: 269-282, 1989. 186. Bai, G. and Weiss, B. The increase of calmodulin in PC12 cells induced by NGF is caused by differential expression of multiple mRNAs for cahnodulin. J. Cell. Physiol., 149: 414-421,1991. 187. Kretsinger, RH. and Nockolds, C.E. Carp muscle calcium-binding protein. 11. Structure determination and general description. J. Biol. Chem, 248: 3313-3326, 1973. 188. Klee, CB. and Vanaman, T.C. Cahnodulin. Adv. Protein. Chem, 35: 213-321, 1982. 189. Means, A.R. Molecular mechanisms of action of calmodulin. Recent. Prog. Horm. Res., 44: 223-262, 1988. 190. Cheung, W.Y. Calmodulin plays a pivotal role in cellular regulation. . Science, 207: 19-27, 1980. 58 191. Means, AR. and Dedman, J .R. Calmodulin--an intracellular calcium receptor. Nature, 285: 73-77, 1980. 192. Cheung, W.Y. Cyclic 3',5'-nucleotide phosphodiesterase. Demonstration of an activator. Biochem. Biophys. Res. Commun., 38: 533-538, 1970. 193. Kakiuchi, S. and Yamazaki, R. Calcium dependent phosphodiesterase activity and its activating factor (PAF) from brain studies on cyclic 3',5'-nucleotide phosphodiesterase (3). Biochem. Biophys. Res. Commun., 41: 1104-1110, 1970. 194. Cheung, W.Y., Bradham, L.S., Lynch, T.J., Lin, Y.M., and Tallant, E.A. Protein activator of cyclic 3':5'-nuclcotide phosphodiesterase of bovine or rat brain also activates its adenylate cyclase. Biochem. Biophys. Res. Commun., 66: 1055-1062, 1975. 195. Brostrom, C.O., Huang, Y.C., Breckenridge, B.M., and Wolff, D.J. Identification of a calcium-binding protein as a calcium-dependent regulator of brain adenylate cyclase. Proc. Natl. Acad. Sci. U. S. A., 72: 64-68, 1975. 196. Gopinath, RM. and Vincenzi, F.F. Phosphodiesterase protein activator mimics red blood cell cytoplasmic activator of (Ca2+-Mg2+)ATPase. Biochem. Biophys. Res. Commun., 77: 1203-1209, 1977. 197. Katz, S. and Remtulla, M.A. Phosphodiesterase protein activator stimulates calcium transport in cardiac microsomal preparations enriched in sarcoplasmic reticulum. . Biochem. Biophys. Res. Commun., 83: 1373-1379, 1978. 198. Biden, T.J., Comte, M., Cox, J .A., and Wollheirn, C.B. Calcium-calmodulin stimulates inositol 1,4,5-trisphosphatc kinase activity from insulin-secreting RINmSF cells. J. Biol. Chem, 262: 9437-9440, 1987. 199. Yamaguchi, K., Hirata, M., and Kuriyarna, H. Calmodulin activates inositol 1,4,5-trisphosphate 3-kinase activity in pig aortic smooth muscle. Biochem. J., 244: 787-791, 1987. 200. Yagi, K., Yazawa, M., Kakiuchi, S., Ohshima, M., and Uenishi, K. Identification of an activator protein for myosin light Chain kinase as the Ca2+-dependent modulator protein. J. Biol. Chem, 253: 1338-1340, 1978. 201. Schulman, H. and Greengard, P. Stimulation of brain membrane protein phosphorylation by calcium and an endogenous heat-stable protein. Nature, 271 : 478-479, 1978. 202. Schulman, H. and Greengard, P. Ca2+-dependent protein phosphorylation system in membranes from various tissues, and its activation by "calcium-dependent regulator". . Proc. 59 Natl. Acad. Sci. U. S. A., 75: 5432-5436, 1978. 203. Wallace, R.W., Lynch, T.J., Tallant, B.A., and Cheung, W.Y. Purification and Characterization of an inhibitor protein of brain adenylate cyclase and cyclic nucleotide phosphodiesterase. J. Biol. Chem, 254: 377-382, 1979. 204. Shanna, R.K., Desai, R., Waisman, D.M., and Wang, J.H. Purification and subunit structure of bovine brain modulator binding protein. J. Biol. Chem, 254 : 4276-4282, 1979. 205. Klee, C.B., Crouch, TH, and Krinks, M.H. Calcineurin: a calcium- and calmodulin-binding protein of the nervous system. Proc. Natl. Acad. Sci. U. S. A., 76: 6270-6273, 1979. 206. Lu, K.P., Rasmussen, CD, May, GS, and Means, A.R. Cooperative regulation of cell proliferation by calcium and calmodulin in Aspergillus nidulans. Mol. Endocrinol., 6: 365-374, 1992. 207. Pelech, S.L. and Sanghera, J .8. MAP kinases: Charting the regulatory pathways [comment]. Science, 25 7: 1355-1356, 1992. 208. Ray, LB. and Sturgill, T.W. Insulin-stimulated microtubule-associated protein kinase is phosphorylated on tyrosine and threonine in vivo. Proc. Natl. Acad. Sci. U. S. A., 85: 3753-3757, 1988. 209. Pelech, S.L. and Sanghera, J .S. Mitogen-activated protein kinases: versatile transducers for cell signaling. Trends. Biochem. Sci., 1 7: 233-23 8, 1992. 210. Victor, 1. and Vilcek, J. Pathways of heat shock protein 28 phosphorylation by TNF in human fibroblasts. Lymphokine. Cytokine. Res., 13: 315-323, 1994. 211. Guy, G.R., Chua, S.P., Wong, N.S., Ng, SB, and Tan, Y.H. Interleukin 1 and tumor necrosis factor activate common multiple protein kinases in human fibroblasts. J. Biol. Chem, 266: 14343-14352, 1991. 212. Wu, J., Michel, H., Dent, P., Haystead, T., Hunt, DE, and Sturgill, T.W. Activation of MAP kinase by a dual specificity Tyr/Thr kinase. Adv. Second. Messenger. Phosphoprotein. Res., 28: 219—225, 1993. 213. Sillirnan, CC. and Sturgill, T.W. Phosphorylation of microtubule-associated protein 2 by MAP kinase primarily involves the projection domain. Biochem. Biophys. Res. Commun., 160: 993-998, 1989. 214. Sanghera, J.S., Aebersold, R., Morrison, H.D., Bures, E.J., and Pelech, S.L. Identification of the sites in myelin basic protein that are phosphorylated by meiosis-activated protein kinase p44mpk. FEBS. Lett., 273: 223-226, 1990. 60 215. Sturgill, T.W., Ray, L.B., Erikson, E., and Maller, J.L. Insulin-stimulated MAP-2 kinase phosphorylates and activates ribosomal protein S6 kinase 11. Nature, 334: 715-718, 1988. 216. Davis, R.J. The mitogen-activated protein kinase signal transduction pathway. . J. Biol. Chem, 268: 14553-14556, 1993. 217. Lin, L.L., Wartmann, M., Lin, A.Y., Knopf, J.L., Seth, A., and Davis, R.J. cPLA2 is phosphorylated and activated by MAP kinase. Cell, 72: 269-278, 1993. 218. Gupta, S. and Davis, R.J. MAP kinase binds to the NH2-terminal activation domain of C-Myc. . FEBS Lett., 353: 281-285, 1994. 219. Seth, A., Gonzalez, F .A., Gupta, S., Raden, D.L., and Davis, R.J. Signal transduction within the nucleus by mitogen-activated protein kinase. J. Biol. Chem, 26 7: 24796-24804, 1992. 220. Davis, R.J. Transcriptional regulation by MAP kinases. Mol. Reprod. Dev., 42: 459-467, 1995. 221. Kahan, C., Seuwen, K., Meloche, S., and Pouyssegur, J. Coordinate, biphasic activation of p44 mitogen-activated protein kinase and S6 kinase by growth factors in hamster fibroblasts. Evidence for thrombin-induced signals different from phosphoinositide turnover and adenylylcyclase inhibition. J. Biol. Chem, 26 7: 13369-13375, 1992. 222. Vouret Craviari, V., Van Obberghen Schilling, E., Scimeca, J.C., Van Obberghen, E., and Pouyssegur, J. Differential activation of p44mapk (ERKI) by alpha-thrombin and thrombin-receptor peptide agonist. Biochem. J., 289: 209-214, 1993. 223. Meloche, S., Seuwen, K., Pages, G., and Pouyssegur, J. Biphasic and synergistic activation of p44mapk (ERKI) by growth factors: correlation between late phase activation and mitogenicity. Mol. Endocrinol., 6: 845-854, 1992. 224. Mansour, S.J., Matten, W.T., Hermann, A.S., Candia, J.M., Rong, S., Fukasawa, K., Vande Woude, G.F., and Ahn, N.G. Transformation of mammalian cells by constitutively active MAP kinase kinase. Science, 265 : 966-970, 1994. 225. Cowley, S., Paterson, H., Kemp, P., and Marshall, C.J. Activation of MAP kinase kinase is necessary and sufficient for PC12 differentiation and for transformation of NIH 3T3 cells. Cell., 77: 841-852, 1994. 226. Pages, G., Lenormand, P., L'Allemain, G., Chambard, J .C., Meloche, S., and Pouyssegur, J. Mitogen-activated protein kinases p42mapk and p44mapk are required for fibroblast proliferation. Proc. Natl. Acad. Sci. U. S. A., 90: 8319-8323, 1993. 61 227. Frost, J.A., Geppcrt, T.D., Cobb, M.H., and Feramisco, J.R. A requirement for extracellular signal-regulated kinase (ERK) function in the activation of AP-l by Ha-Ras, phorbol lZ-myristate 13-acetate, and serum. Proc. Natl. Acad. Sci. U. S. A., 91: 3844-3848, 1994. 228. Marshall, CJ. Specificity of receptor tyrosine kinase signaling: transient versus sustained extracellular signal-regulated kinase activation. Cell., 80: 179-185, 1995. 229. Traverse, S., Gomez, N., Paterson, H., Marshall, C., and Cohen, P. Sustained activation of the mitogen-activated protein (MAP) kinase cascade may be required for differentiation of PC 12 cells. Comparison of the effects of nerve growth factor and epidermal growth factor. Biochem. J., 288: 351-355, 1992. 230. Chao, M.V. Growth factor signaling: where is the specificity? . Cell., 68: 995-997, 1992. 231. Muroya, K., Hattori, S., and Nakamura, S. Nerve growth factor induces rapid accumulation of the GTP-bound form of p21ras in rat pheochromocytoma PC12 cells. Oncogene, 7: 277-281, 1992. 232. Nguyen, T.T., Scimeca, J.C., Filloux, C., Peraldi, P., Carpentier, J.L., and Van Obberghen, E. Co-regulation of the mitogen-activated protein kinase, extracellular signal-regulated kinase 1, and the 90-kDa ribosomal S6 kinase in PC12 cells. Distinct effects of the neurotrophic factor, nerve growth factor, and the mitogenic factor, epidermal growth factor. J. Biol. Chem, 268: 9803-9810, 1993. 233. Traverse, S., Seedorf, K., Paterson, H., Marshall, C.J., Cohen, R, and Ullrich, A. EGF triggers neuronal differentiation of PC12 cells that overexpress the EGF receptor. Curr. Biol., 4: 694-701, 1994. 234. Dikic, 1., Schlessinger, J., and Lax, I. PC12 cells overexpressing the insulin receptor undergo insulin-dependent neuronal differentiation. . Curr. Biol., 4: 702-708, 1994. 235. Denning, M.F., Dlugosz, A.A., Williams, E.K., Szallasi, Z., Blumberg, P.M., and Yuspa, S.H. Specific protein kinase C isozymes mediate the induction of keratinocyte differentiation markers by calcium. Cell Growth Differ., 6: 149-157, 1995. 236. Leboff, M.S., Shoback, D., Brown, B.M., Thatcher, J ., Leombruno, R., Beaudoin, D., Henry, M., Wilson, R., Pallotta, J., Marynick, S., and et a1 , Regulation of parathyroid hormone release and cytosolic calcium by extracellular calcium in dispersed and cultured bovine and pathological human parathyroid cells. J. Clin. Invest, 75 : 49-57, 1985. 237. Shoback, D., Thatcher, J ., Leombruno, R., and Brown, E. Effects of extracellular Ca++ and Mg++ on cytosolic Ca++ and PTH release in dispersed bovine parathyroid cells. Endocrinology, 1 13: 424-426, 1983. 62 238. Brown, E.M., Gardner, D.G., and Aurbach, G.D. Effects of the calcium ionophore A23187 on dispersed bovine parathyroid cells. Endocrinology, 106: 133-138, 1980. 239. Shoback, D.M., Membreno, LA, and MCGhee, J .G. High calcium and other divalent cations increase inositol trisphosphate in bovine parathyroid cells. . Endocrinology, 123: 382-389, 1988. 240. Epstein, P.A., Prentki, M., and Attic, M.F. Modulation of intracellular Ca2+ in the parathyroid cell. Release of Ca2+ from non-mitochondrial pools by inositol trisphosphate. FEBS Lett., 188: 141-144, 1985. 241. Brown, E.M., Gamba, G., Riccardi, D., Lombardi, M., Butters, R., Kifor, 0., Sun, A., Hediger, M.A., Lytton, J ., and Hebert, S.C. Cloning and characterization of an extracellular Ca(2+)-sensing receptor from bovine parathyroid. Nature, 366: 575-580, 1993. 242. Chen, C.J., Barnett, J.V., Congo, DA, and Brown, E.M. Divalent cations suppress 3',5'-adenosine monophosphate accrunulation by stimulating a pertussis toxin-sensitive guanine nucleotide-binding protein in cultured bovine parathyroid cells. Endocrinology, 124: 233-239, 1989. 243. Garrett, J.E., Capuano, I.V., Hammerland, L.G., Hung, B.C., Brown, E.M., Hebert, S.C., Nemeth, ER, and Fuller, F. Molecular Cloning and functional expression of human parathyroid calcium receptor cDNAs. J. Biol. Chem, 270: 12919-12925, 1995. 244. Brown, E.M., Pollak, M., and Hebert, S.C. Sensing of extracellular Ca2+ by parathyroid and kidney cells: Cloning and Characterization of an extracellular Ca(2+)-sensing receptor. Am. J. Kidney. Dis., 25: 506-513, 1995. 245. Brown, E.M., Pollak, M., Chou, Y.H., Seidman, C.E., Seidman, J .G., and Hebert, SC. The cloning of extracellular Ca(2+)-scnsing receptors from parathyroid and kidney: molecular mechanisms of extracellular Ca(2+)-sensing. J. Nutr., 125: 19658-19703, 1995. 246. Brown, E.M., Pollak, M., Chou, Y.H., Seidman, C.E., Seidman, J .G., and Hebert, S.C. Cloning and functional characterization of extracellular Ca(2+)-sensing receptors from parathyroid and kidney. Bone, 17: 7S-1 IS, 1995. 247. Pollak, M.R., Brown, E.M., Chou, Y.H., Hebert, S.C., Marx, S.J., Steinmann, B., Levi, T., Seidman, CE, and Seidman, J .G. Mutations in the human Ca(2+)-sensing receptor gene cause familial hypocalciuric hypercalcemia and neonatal severe hyperparathyroidism [see comments]. Cell, 75: 1297-1303, 1993. 248. Pearce, S.H., Williamson, C., Kifor, 0., Bai, M., Coulthard, M.G., Davies, M., Lewis Bamed, N., McCredie, D., Powell, H., Kendall Taylor, R, Brown, E.M., and Thakker, R.V. A familial syndrome of hypocalcemia with hypercalciuria due to mutations in the 63 calcium-sensing receptor [see comments]. N. Engl. J. Med., 335: 1115-1122, 1996. 249. Chou, Y.H., Pollak, M.R., Brandi, M.L., Toss, G., Amqvist, H., Atkinson, A.B., Papapoulos, S.E., Marx, S., Brown, E.M., Seidman, J.G., and et a1 , Mutations in the human Ca(2+)-sensing-receptor gene that cause familial hypocalciuric hypercalcemia Am. J. Hum. Genet, 56: 1075-1079, 1995. 250. Pearce, S.H., Bai, M., Quinn, S.J., Kifor, 0., Brown, E.M., and Thakker, R.V. Functional Characterization of calcium-sensing receptor mutations expressed in human embryonic kidney cells. . J. Clin. Invest, 98: 1860-1866, 1996. 251. Bai, M., Quinn, 8., Trivedi, S., Kifor, 0., Pearce, S.H.S., Pollak, M.R., Krapcho, K., Hebert, SC, and Brown, E.M. Expression and characterization of inactivating and activating mutations in the human Ca2+o-sensing receptor. J. Biol. Chem, 271: 19537-19545, 1996. 252. Pearce, SH. and Brown, E.M. Calcium-sensing receptor mutations: insights into a structurally and functionally novel receptor [editorial; comment]. J. Clin. Endocrinol. Metab., 81: 1309-1311, 1996. 253. Brown, A.J., Zhong, M., Ritter, C., Brown, E.M., and Slatopolsky, E. Loss of calcium responsiveness in cultured bovine parathyroid cells is associated with decreased calcium receptor expression. Biochem. Biophys. Res. Commun., 212: 861-867, 1995. 254. Mithal, A., Kifor, 0., Kifor, 1., Vassilev, P., Butters, R., Krapcho, K., Sirnin, R., Fuller, F., Hebert, SC, and Brown, E.M. The reduced responsiveness of cultured bovine parathyroid cells to extracellular Ca2+ is associated with marked reduction in the expression of extracellular Ca(2+)-sensing receptor messenger ribonucleic acid and protein. . Endocrinology, 1 3 6: 3087-3092, 1995. 255. Ho, C., Conner, D.A., Pollak, M.R., Ladd, D.J., Kifor, 0., Warren, H.B., Brown, E.M., Seidman, J.G., and Seidman, C.E. A mouse model of human familial hypocalciuric hypercalcemia and neonatal severe hyperparathyroidism [see comments]. Nat. Genet, 11: 389-394, 1995. 256. Takaichi, K. and Kurokawa, K. Inhibitory guanosine triphosphate-binding protein-mediated regulation of vasopressin action in isolated single medullary tubules of mouse kidney. J. Clin. Invest, 82: 1437-1444, 1988. 257. Juhlin, C., Lundgren, S., Johansson, H., Lorentzen, J ., Rask, L., Larsson, E., Rastad, J., Akerstrom, G., and Klareskog, L. SOO-Kilodalton calcium sensor regulating cytoplasmic Ca2+ in cytotrophoblast cells of human placenta. J. Biol. Chem, 265 : 8275-8279, 1990. 258. Juhlin, C., Johansson, H., Holmdahl, R., Gylfe, E., Larsson, R., Rastad, J., Akerstrom, G., and Klareskog, L. Monoclonal anti-parathyroid antibodies interfering with a Ca2+-sensor of human parathyroid cells. Biochem. Biophys. Res. Commun., 143: 570-574, 1987. 64 259. Juhlin, C., Holmdahl, R., Johansson, H., Rastad, J ., Akerstrom, G., and Klareskog, L. Monoclonal antibodies with exclusive reactivity against parathyroid cells and tubule cells of the kidney. Proc. Natl. Acad. Sci. U. S. A., 84: 2990-2994, 1987. 260. Lundgren, S., Hjalm, G., Hellman, P., Ek, B., Juhlin, C., Rastad, J., Klareskog, L., Akerstrom, G., and Rask, L. A protein involved in calcium sensing of the human parathyroid and placental cytotrophoblast cells belongs to the LDL-receptor protein superfamily. . Exp. Cell Res., 212: 344-350, 1994. 261. Saito, A., Pietromonaco, 8., L00, A.K., and Farquhar, M.G. Complete Cloning and sequencing of rat gp330/"megalin," a distinctive member of the low density lipoprotein receptor gene family. Proc. Natl. Acad. Sci. U. S. A., 91: 9725-9729, 1994. 262. Hjalm, G., Murray, E., Crumley, G., Harazim, W., Lundgren, S., Onyango, 1., Ek, B., Larsson, M., Juhlin, C., Hellman, R, Davis, H., Akerstrom, G., Rask, L., and Morse, B. Cloning and sequencing of human gp330, a Ca(2+)-binding receptor with potential intracellular signaling properties. Eur. J. Biochem, 239: 132-137, 1996. 263. Patstone, G. and Maher, P. Copper and calcium binding motifs in the extracellular domains of fibroblast growth factor receptors. J. Biol. Chem, 271: 3343-3 346, 1996. 264. McGrath, GM. and Soule, H.D. Calcium regulation of normal human mammary epithelial cell growth in culture. In. Vitro., 20: 652-662, 1984. 265. Black, BL. and Smith, J.E. Regulation of goblet cell differentiation by calcium in embryonic chick intestine. FASEB. J ., 3: 2653-2659, 1989. CHAPTER II Extracellular Ca2+ stimulates the activation of mitogen-activated protein kinase and cell growth in human fibroblasts Shixia Huang, Veronica M. Maher, and J. Justin McCormick" Carcinogenesis Laboratory- Fee Hall Department of Microbiology and Department of Biochemistry The Cancer Center Michigan State University, East Lansing, MI 48824 65 66 ABSTRACT In serum-free medium containing serum replacements but totally lacking in protein growth factors, diploid human fibroblasts remained quiescent if the extracellular Ca“ concentration was only 0.1 mM. However, when the Ca2+ concentration in this medium was increased to 1 mM, the cells replicated as rapidly as they do in medium supplemented with protein growth factors. When quiescent cells in medium with only 0.1 mM Ca2+ were exposed to 1 mM or 10 mM Ca2+ or 100 ng/ml EGF, the 42 kDa and 44 kDa forms of MAPK were rapidly activated, as demonstrated by a Characteristic electrophoretic mobility shift of these proteins and by their enhanced ability to phosphorylate myelin basic protein (MBP). Analysis of fractions from Mono Q anion-exchange Chromatography of lysates of cells exposed to 10 mM Ca2+ or 100 ng/ml EGF revealed a peak of MBP phosphorylation activity that coeluted with p42 and p44 MAPK as shown by immunoblot analysis. Activation of MAPK by extracellular Ca2+ was dose dependent and biphasic, with a peak of activation at 5-10 min after exposure, followed by a period of sustained activation of MAPK at a lower level. This pattern has been shown (Vouret-Craviari et al. 1993, Biochem J. 289, 209) to correlate with the re-entry of mammalian cells into the cell cycle. 67 ABBREVIATIONS EGF, epidermal growth factor; HRP, horseradish peroxidase; MAPK, mitogen- activated protein kinase; MBP, myelin basic protein; SCS, supplemented calf serum. Running Title: Extracellular Ca2+ stimulates MAPK and cell growth 68 INTRODUCTION It is known that human fibroblasts in culture will replicate in medium containing specific protein growth factors (1), and that Ca2+ is an important factor in cell metabolism (2). Studies show that such protein growth factors as EGF, platelet-derived growth factor, insulin, thrombin, nerve growth factor, tumor necrosis factor, and bombesin, can initiate a phosphorylation cascade in various cell types, including human fibroblasts, and that one of the key events of this cascade is the activation of MAPK by phosphorylation (3-9). Several closely related MAPK isoforms, referred to by their molecular masses, i.e., p40, p42, p44, and p54 (10), have been identified in various cell types. In human fibroblasts, EGF (5), tumor necrosis factor (7,] l) and interleukin (1]) have been shown to cause activation of p42 and/or p44 MAPKs. Activated MAPK can, in turn, phosphorylate MBP, microtubule- associated protein 2, the EGF receptor, phospholipase A2, and the nuclear transcriptional factors c-myc, C-fos, C-jun (10, 12, 13). Morgan et al. (14) in this laboratory showed that diploid human fibroblasts can replicate rapidly in serum-free MCM medium (15), a modified version of MCDB 110 (1), if supplied with the serum replacements, e.g., lipids, iron, attachment factors, ete, specified by Ryan et al. (15) and with insulin as the only protein growth factor, instead of the two growth factors specified by Ryan et al. (15), e.g., insulin and EGF. MCM medium, as ordinarily prepared (15), contains 1 mM Ca2+. Morgan et al. (14) also showed that the cells do not replicate if the Ca2+ concentration is reduced from 1 mM to 0.1 mM, suggesting that 69 extracellular Ca2+ stimulates cell growth. We tested this hypothesis by removing all protein growth factors from the medium and seeing if Ca2+ alone could cause the cells to replicate for an extended period. Since MAPK is known to integrate the signals from various growth factor pathways, we also tested whether Ca2+ causes the activation of MAPK in these cells as protein growth factors do. The answer to both questions was positive. In the total absence of protein growth factors, extracellular Ca2+ supported growth of the cells during the two weeks of the experiment. When quiescent cells were stimulated with Ca2+ or EGF, MAPK activation was observed as demonstrated by a characteristic shift in the electrophoretic mobility of MAPK and an enhanced ability of MAPK to phosphorylate MBP. 70 MATERIALS AND METHODS Materials MBP, EGF, and B-glycerophosphate were purchased from Sigma (St. Louis, MO). Polyclonal anti-MAPK antibody Ab283 (from rabbit) was kindly provided by Dr. Marsha R. Rosner of the University of Chicago. Monoclonal anti-MAPK antibodies were purchased from Chemicon (Temecula, CA) and Zymed Laboratories (South San Francisco, CA). HRP- linked goat anti-mouse IgG was obtained from BioRad Laboratories (Richmond, CA) and HRP-conjugated goat anti-rabbit IgG was from Boehringer Mannheim Corp. (Indianapolis, IN). Cells and cell culture Diploid human fibroblast cell lines LG] and SL80, derived in this laboratory fi'om neonatal foreskins, were used at cell population doubling 15 to 28. The in vitro life span of LG] cells is 40 population doublings, that of SL80 cells is 80 population doublings. Cells were routinely grown in Eagle's minimum essential medium supplemented with 0.2 mM aspartic acid, 1.0 mM sodium pyruvate, and 0.2 mM serine, 10% supplemented calf serum (SCS) (Hyclone Laboratories, Logan UT) penicillin (100 units/ml), streptomycin (100 ug/ml), and hydrocortisone (1 ug/ml) (complete medium) at 37°C in humidified incubator with 5% (30,. Media for testing for growth stimulation 71 To assay for growth stimulation we used MCM medium (15), a modified version of MCDB 110 (1) which was developed for the continuous proliferation of human fibroblasts. We used the serum replacements specified by Ryan et al. (15) but omitting the protein growth factors, i.e., EGF and insulin. The medium was prepared to contain 0.1 mM or 1 mM Ca2+ as indicated. Preparation of cell lysates Cells were grown in 100-mm diameter dishes in complete medium until subconfluent or confluent as desired. To make the cells quiescent, the medium was replaced for 24to 30 h with MCM medium containing serum replacements but no protein growth factors and with the Ca2+ concentration reduced to 0.1 mM or to zero where indicated. Unless otherwise indicated, increased Ca2+ or EGF was then added directly to the medium on the cells with a gentle swirling, and the incubation at 37°C was continued for the indicated time. To harvest the cells, the medium was removed, the cell sheet was quickly washed twice with Ca2+-free, Mg“-free PBS, and the cells were lysed in 0.3 to 0.5 ml of buffer composed of 20 mM Tris- HCl, pH 8.0, 137 mM NaCl, 1% Nonidet P-40, 10% glycerol, and 1 mM phenylrnethylsulphonylfluoride, 5 mM EDTA, 0.15 U/ml aprotinin, and 1 mM sodium orthovanadate. The cells were collected with a rubber scraper and kept on ice for 20 min. The lysates were centrifuged for 10 min at 4°C in a microfuge, and the protein concentration of the supernatant was determined using a Bicinchroninic Acid Protein Assay kit (Pierce Chemical Co., Rockford, IL) with BSA as the protein standard. The cell lysates were used immediately after preparation or stored at -80°C until used. 72 Immunoblot Analysis of MAPK Cell lysates containing 25 ug of protein were dissolved in buffer containing 0.05 M Tris-HCI, pH 6.9, 0.72 M mercaptoethanol, 9% glycerol, 2.3% SDS and 0.1% Bromphenol blue (sample buffer), and the proteins were separated by 10% SDS-PAGE, and transferred from the gels onto Immobilon PVDF transfer membranes (Millipore Corp., Bedford, MA) overnight at 35 mA, and the blots were blocked by incubating for 1 h at room temperature in Tris-buffered saline, pH 7.6, (20 mM Tris-HCl, 137 mM NaCl) containing 0.1% Tween 20 (v/v) and 5% nonfat dry milk (w/v). To test for MAPK, the membranes were incubated for l h with the indicated anti-MAPK antibodies in Tris-buffered saline containing 0.1% Tween 20 and 5% nonfat dry milk, washed several times in this Tris-buffered saline, and then incubated with goat anti-rabbit IgG or goat anti-mouse IgG as desired. Enhanced chemiluminescence (Amersham, Arlington Heights, IL) was used as a detection system according to the manufacturer's instructions. MBP phosphorylation as an assay for MAPK activity This was assayed essentially as described previously (16). Briefly, cell lysates containing 2.5 ug of protein were incubated with MBP at 0.5 mg/ml for 10 min at room temperature in a final volume of 20 ul containing 18 mM HEPES, pH 7.4, 10 mM magnesium acetate, and 2 uCi [7-32P]ATP (2-10 Ci/mmol) (DuPont, New England Nuclear Research Products, Boston, MA). The reaction was stopped by the addition of 5X sample 73 buffer. After 5 min at 95°C, proteins were separated on a 10% SDS-PAGE. The dried gel was used to expose Kodak X-Omat film and a phosphor screen. The phosphor screen was scanned and the amount of MBP phosphorylation was quantified using a Phosphorlmager (Model 400B, Molecular Dynamics, Sunnyvale, CA) according to the manufacturer's instructions. Purification of MAPK by mono Q column chromatography Cell lysates from Ca2+-stimulated or EGF-stimulated cells or untreated cells were prepared and normalized for protein concentration. MAPK was purified by Mono Q anion- exchange Chromatography as described (17). Before purification, lysates were thawed and centrifuged for 5 min in a microfuge, and the supernatant containing 1.5 to 2 mg of protein was diluted 2-fold in column buffer, which contains 50 mM B-glycerophosphate, pH 7.2, 100 M sodium orthovanadate, 1 mM EGTA, and 1 mM dithiothreitol. The diluted sample was filtered through a 0.2 pm filter and loaded onto a Mono Q HRS/5 column FPLC (Pharmacia, Piscataway, NJ) (1.5 ml bed volume) at a flow rate of 0.4 ml/min. Using a linear gradient of 0 to 0.4 M NaCl to elute, 1 ml fiactions were collected. Aliquots of 10.5 it] were assayed for MAPK activity using the MBP phosphorylation assay, and aliquots of 110 pl were assayed by immunoblot analysis, as described above. 74 RESULTS AND DISCUSSION Evidence that extracellular Ca2+ stimulates cell growth To determine if extracellular Ca2+ could stimulate diploid human fibroblasts to replicate in the total absence of protein growth factors, we plated LG] and SL80 cells in MCM medium containing 0.1 mM Ca2+ and 1% SCS. The next day, the number of attached cells was determined, and the medium was replaced with MCM medium containing the serum replacements of Ryan et al. (15) but minus the protein growth factors. The rate of replication of the cells in this medium supplemented with Ca2+ or EGF or 10% SCS is shown in Figure 1. Under the most stringent condition, 0.1 mM Ca2+, the number of cells increased only slightly or at most doubled. In the presence of 1 mM Ca2+, the LG] cells doubled every 48 h, and the SL80 cells every 36 h. This rate of growth was as fast or faster than the cells' rate of growth in medium containing EGF (10 ng/ml), and in the case of SL80 cells, almost as fast as their rate with 10% SCS. Higher concentrations of EGF do not increase the growth rate, these results indicate that 1 mM Ca2+ stimulates human fibroblasts to proliferate as they do in response to protein growth factors. MAPK activation as a result of Ca2+ or EGF stimulation To determine whether MAPK is involved in activation of the signal pathway induced by extracellular Ca2+ in human fibroblasts, we treated LG] and SL80 cells for 5 min with 1 mM Ca”, 10 mM Ca2+, or 100 ng/ml EGF, and assayed the cell lysates for evidence of MAPK activation using immunoblot analysis. Protein growth factors are typically used at 75 I SL80 106 106‘; ‘///‘ a 3 V m E, 105 105‘ V/ 4 LARA IPT'II 68101 2468101214 2 Days 214 #- Figure 1 Growth stimulating activity of Ca’+ Cells were plated into 60 mm-diameter dishes in MCM medium containing 0.1 mM Ca" and 1% SCS. The next day (day 1), the number of attached cells was determined, the medium was replaced with MCM medium containing the serum replacements of Ryan et al. (15) minus insulin and EGF, but modified as follows: 0.1 mM Ca2+ (O); 1 mM Ca” (0); 0.1 mM Ca” with 10 ng/ml EGF (V); 0.1 mM Ca” with 10% SCS (V). The cells were refed with appropriate medium on day 4, 7, and 10, and the number of cells in four dishes each time for each condition was determined on day 4, 7, 10 and 14. Similar results were obtained in two separate experiments. 76 ~10 higher concentrations in MAPK assays than in growth studies. We used anti-MAPK polyclonal antibody Ab283 which was previously shown to detect both the 42 and 44 kDa isoforms of MAPK (5). With the cell lysates prepared from untreated cells, this antibody revealed two bands with molecular masses of approximately 42 and 44 kDa (lane 1 of Figure 2A and ZB). With cell lysates prepared from cells exposed for 5 min to 1 mM or 10 mM Ca2+ or to 100 ng/ml EGF (lanes 2, 3, and 4 of Figure 2A and 2B), slower moving bands were observed, one just above the 42 kDa band and one just above the 44 kDa band. This molecular weight shift is Characteristic of the phosphorylated form of the MAPK isoforms (18, 19). The percentage of each MAPK isoform that traveled more slowly was greater with lysate from cells exposed to 10 mM Ca2+ or IOOng/ml EGF than from cells exposed to 1 mM Ca2+. We also stripped the membrane using 62.5 mM Tris-HCI, pH 6.9, 2% SDS, and 0.1 M mercaptoethanol for 30 min at 50°C and reprobed it with monoclonal antibodies that recognize only the 42 kDa MAPK in human cell lines. The results indicated that the lower bands are the p42 form of MAPK (data not shown). To determine whether this phosphorylation of MAPK resulted in a molecule with kinase activity, the cell lysates were also tested as described for their ability to phosphorylate MBP . As shown in Figure 2C and 2D, Ca2+ or EGF increased the degree of MBP phosphorylation. The degree of MBP phosphorylation produced by lysates from cells stimulated with 10 mM Ca2+ was much greater than that from cells given 1 mM Ca2+, and approximately equal to that from the cells treated with lOOng/ml EGF. The activation of MAPK by Ca“ or lOOng/ml EGF was also evaluated on Mono Q column-purified MAPK from the cell lysates. The lysates from untreated cells or from cells 77 .m~:oEtoexc concern: beg—058:5 00:: :8 033E323: 2: m_ :32? 3.23: 2.; .mOm _E\w: oo— .e 2:: {mu SE 2 .m 33 5:0 2:: _ .N .33 “Leo E _.o .0: 405:8 .\ 3:3 62:03: m: 00.33: 0:: .EE o. :8 93.32-: :5 mm: 53» BEE 203 3:85 :33 moEEam «.28». :8 .3 n52 e0 :0_a_b0:em0:e ”Q U gage .mmmn< E2335 VE< SEE: 5:5 @305 warns: ~0_n0::EE_ 2 033.33 E: .m O01 -‘ N (A O O O L 1 lg O 0 2 4 6 8101214161820 SL80 O O in N00! ANN-#01 COO MBP phCSphorylotion(counts x 10—5) 0 O O 2 4 6 8101214161820 Fraction Figure 3 Assay of mono Q fractions for ability to phosphorylate MBP Cell lysates were prepared as described from cells exposed to 10 mM Ca2+ (O ) or 100 rig/ml EGF (V) for 5 min or left untreated (O). Lysate protein (1.5 to 2 mg) was diluted by column bufl‘er and loaded onto a Mono Q column. The protein was eluted by a linear NaCl gradient, and 1 ml fractions were collected. Aliquots of fraction 8 through 20 were assayed for MBP phosphorylation. The salt gradient is denoted by the broken line. Similar results were obtained in two experiments. 80 955598 025 :0: 0303:0825: 05 m_ 950% $38: 2:. 30000:: MAE—2-0:: 505 009580 3 0000.5 0:: mares: 830::EE_ 9 0800.33 203 80:25 082380 05 e0 30:32. .m 2:3 “— E 550.0 E 005800 203 20008: 0 0:02 20:00.... 0 0:02 «0 Emu—«:0 .030:=EE— v 0.59...— ‘ U - 1‘ -' U ‘ «on ' ale ‘ E. 1.... 00a ’0 up 3 2. 3 9 up up up .02 an. e O «0» 00 600 (ID? x «[60 14.. I09. 03m ’04 8l circumstances, when 10 mM Ca2+ was added to the solution on the cells, MAPK activation was observed (data not shown). To determine if there was a dose-response relationship for Ca2+- induced MAPK activation, quiescent LGl and SL80 cells in McM medium lacking Ca2+ and supplemented with serum replacements, but without protein growth factors, were exposed for 5 min to various concentrations of Ca”. Cell lysates were prepared and analyzed for their ability to phosphorylate MBP. A dose response was observed with both cell lines (Figure 5). At high concentrations of Ca2+, MAPK activation would be expected to exhibit saturation; however, at concentrations of 50 mM Ca2+ or higher, a precipitate formed, making it impossible to extend the curve. Time course for Ca2+-induced MAPK activation Pouysse'gur and his colleagues showed that in quiescent Chinese hamster fibroblast (cell line CCL39), MAPK is activated by thrombin or basic fibroblast growth factor and that this activation exhibits a biphasic pattern i.e., a rapid increase in activity followed by a rapid decrease and then a sustained level of lower activity (22, 23). They proposed that this sustained activity is required for the re-entry of these cells from G0 into the cell cycle (23), i.e., for triggering the proliferative response, because when they treated CCL39 cells with a thrombin analog instead of thrombin or with carbachol, the second phase of MAPK activation was not observed and no DNA synthesis took place. To see if Ca2+-stimulated MAPK activation in human fibroblasts also exhibits such a biphasic pattern, we carried out a time course for Ca2+ stimulation, using phosphorylation of MBP as a measure of MAPK 82 (I) O O O) O O 1 C d N O O .1 T” "//f I I I I I I 1 o 10‘1 10° 101 102 Calcium (mM) MBP phosphorylation (percent of control) 8 o C Figure 5 Dose-response of Ca2+-stimulated MAPK activity Cells were grown to confluenceand the medium was changed to McM medium with serum replacements of Ryan et al. (15) but no protein growth factors, and modified to be without Ca”. After 24 h, the indicated concentration of Ca2+ was applied. After 5 min, cell lysates were prepared and assayed for MBP phosphorylation. Similar results were obtained in two separate experiments. 83 activation. The results are shown in Figure 6. Exposure of LG] and SL80 cells to 10 mM Ca2+ rapidly activated MAPK, with a maximum peak of activation occurring at 5 to 10 min. Afier the initial burst of activation, a second wave of sustained activation was observed in both cell lines. The level of activation of MAPK did not come down to the basal level in either cell line during the 180 min of stimulation. This sustained activation, along with our finding that in the total absence of protein growth factors, Ca2+ induces growth of human fibroblasts (Figure 1), suggests that Ca2+ stimulates cell proliferation by activating MAPK in a biphasic manner. Recently, Frost et al. (24) and Pages et al. (25), using MAPK antisense RNA and/or MAPK kinase-deficient mutants to suppress MAPK activation, showed that MAPK activation is required for activation of transcriptional factors (24) and for rodent fibroblast proliferation (25). In summary, as far as we can determine, our study is the first to show that in the total absence of protein growth factors extracellular Ca2+ can stimulate sustained growth of diploid human fibroblasts and can induce the activation of MAPK. In the absence of protein growth factors, the 42 kDa and 44 kDa isoforms of MAPK were both activated, as they are by protein growth factors. A study by Chao et al (5) indicates that various pathways can lead to MAPK activation in human fibroblasts. For example, EGF-mediated activation does not require intracellular Ca2+, whereas thapsigargin-mediated activation depends on intracellular Ca2+. We intend to determine whether extracellular Ca2+ stimulates MAPK activation by one of these pathways or by some alternative mechanism. 84 800 oLG1 oSLBO 600 400 200 \ MBP phosphorylation (percent of controD O I I I I I r I I O 50 100 150 200 Exposure tune (nun) Figure 6 Time course for Ca’*-induced MAPK activity The medium on cells in exponential growth was changed to the MCM medium with serum replacements of Ryan et al. (15) but no protein growth factors and only 0.1 mM Ca”. Afier 24 to 30 h, the Ca2+ concentration was increased to 10 mM for the indicated time. Cell lysates were prepared and assayed for MBP phosphorylation. Results are expressed as means 2’: S.E.M. of triplicate determinations. Similar results were obtained in two separate experiments. 85 ACKNOWLEDGEMENTS We thank Dr. Marsha. R. Rosner for kindly supplying us with the anti-MAPK antibody and Dr. Robert P. Hausinger of Michigan State University for his advice on Mono Q purification of MAPK. This research was supported in part by DHHS Grant AG11026 from the National Institute on Aging and DOE Grant ER60524. 10. 11. 12. 13. 14. 15. 16. 17. 86 REFERENCES Bettger, W. J., Boyce, S. T., Walthall, B. J ., and Ham, R. G. (1981) Proc. Natl. Acad. Sci. USA78, 5588-592 Brown, E. M. (1991) Physiol. Rev. 71, 371-411 Ahn, N. G., Seger, R., Bratlien, R. L., Diltz, C. D., Tonks, N. K., and Krebs, E. G., (1991) J. Biol. Chem. 266, 4220-4227 Nakielny, S., Cohen, P., Wu, J ., and Sturgill, T. (1992) EMBO J. 11, 2123-2129 Chao, T. 0., Byron, K. L., Lee, K. M, Villereal, M., and Rosner, M, R. (1992) J. Biol. Chem. 267, 19876-19883 Pelech, S. L., and Sanghera, J. S. (1992) Science 257, 1355-1356 Victor, 1., Schwenger, R, Li, W., Schlessinger, J., and Vilcek, J. (1993) J. Biol. Chem. 268,18994-1 8999 Pang, L., Decker, s. J., and Saltiel, A. R. (1993) Biochem. J. 289, 283-287 Burgering, B. M. T., de Vries-Smits, A. M. M., Medema, R. H., van Weeren, P. C., Tertoolen, L. G. J., and Bos, J. L. (1993) Mol. Cell. Biol. 13, 7248-7256 Pelech, S. L., and Sanghera, J. S. (1992) TIBS 17, 233-238 Guy, G. R., Chua, S. P., Wong, N. S., Ng, S. B., and Tan, Y. H. (1991) J. Biol Chem. 266, 14343-14352 Sturgill, T. W., Ray, L. B., Erikson, E., and Maller, J. L. (1988) Nature 334, 715-718 Davis, R. J. (1993) J. Biol. Chem. 268, 14553-14556 Morgan, T. L., Yang, D., Fry, D. G., Hurlin, P. J., Kohler, S. K., Maher, V. M., and McCormick, J. J . (1991) Exp. Cell Res. 197, 125-136. Ryan, P. A., McCormick, J. J., and Maher, V. M. (1987) Exp. Cell Res. 172, 318-328 Zheng, C. F., and Guan, K. L. (1993) J. Biol. Chem. 268, 11435-11439 Gupta, S. K., Gallego, C., Johnson, G. L., and Heasley, LE. (1992) J. Biol. Chem 267, 18. 19. 20. 21. 22. 23. 24. 25. 87 7987- 7990 Welham, M. J., Duronio, V., Sanghera, J. S., Pelech, S. L., and Schrader, J. W. (1992) J. Immunol. 149, 1683-1693 Posada, J ., and Cooper, J. A. (1992) Science 255, 212-215 Offermanns, S., Jones, 8. V. P., Bombien, E., and Schultz, G. (1994) J. Immunol. 152, 250-261 Boulton, T. G., Nye, S. H., Robbins, D. J ., Ip, N. Y., Radziejewska, E., Morgenbesser, S. D., Depinho, R. A., Panayotatos, N., Cobb, M. H., and Yancopoulos, G. D. (1991) Cell 65, 663-675 Kahan, C., Seuwen, K., Meloche, S., and Pouyssegur, J. (1992) J. Biol.Chem. 267, 13369-13375 Vouret-Craviari, V.,Van Obberghen-Schilling, E., Scimeca, J. C., Van Obberghen, E., and Pouyssegur, J. (1993) Biochem. J ., 289, 209-214 Frost, J. A., Geppcrt, T. D., Cobb, M. H., and Feramisco, J. R. (1994) Proc. Natl. Acad. Sci. USA 91, 3844-3848 Pages, G., Lenormand, P., L'Allemain, G., Chambard, J. C., Meloche, S., and Pouyssegur, J. (1993) Proc. Natl. Acad. Sci. USA 90, 8319-8323 CHAPTER III Characterization of the second messenger pathway of extracellular Ca2+-induced mitogen-activated protein kinase activation in human fibroblasts Shixia Huang, Veronica M. Maher, and J. Justin McCormick“ Carcinogenesis Laboratory- Fee Hall Department of Microbiology and Department of Biochemistry The Cancer Center Michigan State University, East Lansing, MI 48824 88 89 ABSTRACT Human fibroblasts in culture will grow in serum free medium containing serum replacement factors but without protein growth factors as long as the Ca2+ level is 1.0-2.0 mM. When the Ca2+ is reduced to 0.1 mM, the cells stop cycling but can be reinduced to cycle by raising the Ca2+ level. Mitogen-activated protein kinase (MAPK) is activated in this process just as it is by protein growth factors (Huang et al. [1995] Biochemical J. 310, 881- 885). We now report that the exposure of human fibroblasts to extracellular Ca” increased cytosolic inositol (1, 4, 5)-trisphosphate level and caused a transient rise in the concentration of intracellular Ca2+. Ca2+-induced MAPK activation was partially abolished by treatment of the cells with pertussis toxin which inhibits certain species of G proteins. It was also decreased by treatment with thapsigargin which depletes intracellular Ca2+ stores, by exposure to phorbol 12-myristyl l3-acetate that downregulates protein kinase C (PKC), by treatment with calmodulin antagonists, N-(6-aminohexyl)-5-chloro-1- naphthalenesulfonamide HCL and calmidazolium, as well as by exposure of the cells to lanthanum, a Ca2+ channel inhibitor. We found that the c-raf-l protein was not phosphorylated after Ca2+ stimulation, and the Ca2+ sensing receptor (Brown et al. [1993] Nature 366, 575-5 80) was not involved in mediating MAPK activation or cell growth. These results suggest that extracellular Ca2+ stimulates MAPK activation through a pathway(s) involving a pertussis toxin sensitive G protein, phospholipase C, intracellular Ca2+, calmodulin, and PKC. 90 ABBREVIATIONS DAG, diacylglycerol; EGF, epidermal growth factor; F BS, fetal bovine serum; 1P3, inositol (1, 4, 5)-trisphosphate; MAP-2, microtubule-associated protein 2; MAPK, mitogen- activated protein kinase; PDGF, platelet-derived grth factor; PIP2, phosphatidylinositol bisphosphate; PLC, phospholipase C; PKC, protein kinase C; PMA, phorbol 12-myristyl 13-acetate; SCS, supplemented calf serum; W-7, N-(6-aminohexyl)-5-chloro-l- naphthalenesulfonamide HCL; W-12, N-(4-aminobutyl)-2-naphthalenesulfonamide. 91 INTRODUCTION It is well known that the extracellular Ca2+ concentration can regulate hormonal secretion [1] and cell differentiation [2-4] in some cell types. For example, mouse and human keratinocyte cells in culture [3-5] proliferate in a medium containing Ca2+ which is 5-40 times (0.05-0.1 mM) lower than the physiological level (1 .0-2.0 mM), but when the Ca2+ concentration of the medium is increase to the physiological range, these cells undergo terminal differentiation. Exposure of these cells to medium with physiological levels of Ca2+ causes a rapid increase in phosphoinositide turnover and intracellular Ca2+ levels [2,3] and the phosphorylation of tyrosine in various proteins [4]. Mesenchymal cells exhibit the opposite response. Normal human [6], mouse [7], and chicken [8-10] fibroblasts in culture stop proliferating when placed in a medium with reduced Ca2+ (0.01-O.1 mM). However, when the Ca2+ concentration is increased to the physiological level, the cells proliferate. We recently showed [11] that when one exposes human fibroblasts to medium containing Ca2+ at physiological levels of or higher but lacking protein growth factors, mitogen-activated protein kinase (MAPK) activation is induced and this correlates with cell growth. In mouse 3T3 fibroblasts in culture, it has been shown that extracelluar Ca2+ stimulates DNA synthesis [12-17] and c-fos expression [14-16,18] similar to that induced by platelet-derived growth factor (PDGF) [15,16]. Furthermore, Epstein et al [15,16] showed that both PDGF-induced and Ca2 + -induced c-fos expression is dependent on protein kinase C (PKC). However, Ca2+ does not induce the phosphorylation of the PDGF receptor and the raf protein as PDGF does, indicating that the pathways which result in similar gene 92 transcription are at least in part independent. As is well known, protein growth factors, such as epidermal growth factor (EGF), PDGF, insulin, thrombin, nerve growth factor, and bombesin, bind to specific receptors on the external cell surface, and initiate a phosphorylation cascade through the activation of either tyrosine kinase-linked or G protein-coupled receptors [19-25]. One of the key events in this cascade is the activation of MAPK by phosphorylation. Two of the well understood pathways leading to MAPK activation from various growth factor stimulation are the ras/raf pathway [19], and phospholipase C (PLC) pathway [26,27]. The activation of PLC, by either type of activated receptor, causes hydrolysis of the inositol lipid-phosphatidylinositol bisphosphate (PIP2), and generates the intracellular second messengers [26,27], diacylglycerol (DAG), and inositol (1, 4, 5)-trisphosphate (1P3). DAG is responsible for activating PKC isoenzymes [28], which leads to the activation of many other kinases including c-raf [29,30] and MAPK [31]. 1P3 binds to the 1P3 receptor on the endoplasmic reticulum to release Ca2+ fi'om intracellular stores [27], and the free Ca2+ is responsible for the induction of many cellular events, including MAPK activation [21]. Intracellular Ca2+ exerts its effects by binding to proteins such as calmodulin [32]. How human fibroblasts “sense” the concentration of extracellular Ca2+ is unknown. One possibility is that there is some type of receptor on the external cell membrane analogous to the receptors for protein growth factors. Recently, Brown et al. [33] cloned a Ca2+-sensing receptor from bovine parathyroid [33], and subsequently, the human homolog of that receptor was cloned [34]. This receptor, which couples to G protein (s), has been shown to mediate the release of the extracellular Ca2+-regulating parathyroid hormone. Once the receptor is stimulated by extracellular Ca”, it increases the intracellular level of Ca2+ and 93 1P3 [33]. A knockout mouse lacking this receptor has recently been generated by Ho et al [35]. Homozygous Ca2+ receptor-deficient mice exhibit severe neonatal hyperparathyroidism, markedly elevated serum levels of Ca2+ and parathyroid hormone, parathyroid hyperplasia, bone abnormalities, and retarded growth [35]. Our interest was in determining how exposure to physiological levels of Ca2+ is able to cause the growth of human fibroblasts in culture. This is of particular importance since, unlike their normal counterparts, human fibrosarcoma-derived fibroblasts [36] and human fibroblasts malignantly transformed in culture [37,3 8] are able to grow in medium with a reduced Ca2+ concentration, even in the absence of protein growth factors. This suggests that some aberration in this pathway plays a causal role in the transformation of these cells. Since exposure of mouse fibroblasts to high concentrations of extracellular Ca” stimulates c-fos transcription via PKC [15], we first examined the possible involvement of PKC in Ca2+ -induced MAPK activation in human fibroblasts. Intracellular Ca2 * measurements were also carried out because studies showed that extracellular Ca2+ regulates intracellular Ca2+ elevation in other cell system [1-3]. Our finding evidence of the involvement of PKC and intracellular Ca2+ led us to examine their downstream effectors c- raf-l and calmodulin respectively, and their upstream activators, PLC, G protein, and the possible involvement of the Ca2+ sensing receptor. The results showed that the addition of extracellular Ca” activated MAPK through a pertussis toxin-sensitive G protein, and led to the elevation of intracellular 1P3 and Ca2+. The activation of MAPK by extracellular Ca2+ is dependent on PKC, intracellular Ca2+, and calmodulin, but is not dependent on c-raf-l , or the Ca2+-sensing receptor we examined. 94 MATERIALS AND METHODS Materials EGF, PDGF-BB, lys-bradykinin, thapsigargin, N-(6-aminohexyl)-5-chloro-1- naphthalenesulfonamide HCL (W-7), N-(4-aminobutyl)—2-naphthalenesulfonamide (W-12), calmidazolium (compound R24571), nifedipine, verapamil, lanthanum chloride were purchased from Sigma (St. Louis, MO, USA). Pertussis toxin was from Calbiolchem (La Jolla, CA, USA). Phorbol 12-myristyl l3-acetate (PMA) (from Calbiolchem) was a gift from Dr. David DeWitt of Michigan State University. Indo-l-Am, cell permeate, and indo-l, cell irnperrneate, were from Molecular Probe (Eugene, OR, USA). Ionomycin (fi'ee acid) and the 1P3 Biotrak Radioirnmunoassay System Kit (TRK 1000) were from Amersham Life Science Inc. (Arlington Heights, IL, USA). Lab-tek, 4-well chambered coverglass (Nuct#l36420) was from VWR Scientific (Chicago, IL, USA). Monoclonal anti-MAPK antibodies were purchased from Chemicon(Temecu1a, CA, USA) and Zymed Laboratories (South San Francisco, CA, USA). Monoclonal anti-c-raf-l antibody was from Transduction Laboratories (Lexington, KY, USA). Horseradish peroxidase-linked goat anti-mouse IgG was obtained from Bio-Rad Laboratories (Richmond, CA, USA). Supplemented calf serum (SCS) and fetal bovine serum (FBS) were from Hyclone Laboratories (Logan, UT, USA). Cells and cell culture Diploid human fibroblast cell line LGl, derived in this laboratory from neonatal foreskin, was used in this study. Cells were routinely grown in Eagle's minimum essential medium supplemented with 0.2 mM aspartic acid, 1.0 mM sodium pyruvate, and 0.2 mM 95 serine, 10% supplemented calf serum (SCS), penicillin (100 units/ml), and streptomycin (100 ug/ml) (complete medium) at 37°C in an humidified incubator containing 5% C02. For studies on the effects of various growth conditions, medium on confluent or subconfluent cells was changed to modified McM medium, a derivative of MCDB-110 medium [39], made with 0.1 mM Ca2+ and supplemented with serum replacements [40] but without any protein growth factors [1 1]. This medium will be referred as serum free medium. Isolation and culture of mouse skin fibroblasts Pregnant female mice (with Ca2+-sensing receptor status of +/- that had previously been mated with +/- male mice) were kindly provided by Dr. Christine Seidman of Harvard Medical School [3 5]. When the litter was born, newborn to 4 day-old mice were sacrificed with CO2 and washed with 70% ethanol. The skin was removed to a 60 mm dish and washed twice in PBS, then transferred to a 50 ml sterile glass flask and minced finely with scissors. Ten ml of 0.24% trypsin was added to the flask containing the mouse skin, and the tissue was incubated at 37°C for 30 min to 1 hour, and shaken every 5 min. The suspension was then centrifuged at 400xg for 4 min to remove tissue fragments, and the turbid supernatant containing single cells was transferred to a 15 ml tube containing 4 ml Eagle’s medium with 20% SCS and centrifuged at 2000xg for 10 min. The pellet was suspended in 3 ml Eagles medium containing 20% F BS, penicillin (600 units/ml), and streptomycin (600 ug/ml). To prevent possible bacteria contamination, a concentration of penicillin and streptomycin six times higher than that normally used in cell culture medium was used. Cells were counted and plated at 4 x 105 to 8 x 105 per 100 mm dish in the same medium, and incubated at 37°C in a humidified incubator containing 5% C02. The next day, medium in each dish was 96 removed, and cells were washed with warm Eagles medium once, and fresh medium, i. e., Eagles medium containing 20% FBS and the noted high concentration of antibiotics, was added into each dish. About 5 days later, confluent cells in each dish were trypsinized and expanded in Eagles medium with 10% FBS, and regular penicillin and streptomycin concentration (100 ug/ml). Media for testing for growth stimulation of mouse fibroblasts The same serum free medium that we employed previously for studying growth stimulation in human fibroblasts [11], was used for mouse skin fibroblasts. With growth stimulation assays, serum free medium was used with no additives, i. e., at 0.1 mM Ca2+, or with 1 mM Ca2+, or supplemented with 10% FBS. Preparation of cell lysates Cells were grown in 60 or 100 mm-diameter dishes in complete medium until confluent. Then the medium was replaced for 24-30 hours with serum free medium. Unless otherwise indicated, increased Ca2+ or EGF or other reagents were added directly to the medium on the cells. Cells treated with different reagents or left unheated were washed twice in cold Ca2+ -free, Mg2+ -fiee PBS, and the cell lysates were prepared as described previously [11]. Immunoblot analysis of MAPK (Western Blot) Cell lysates containing 25 ug of protein were dissolved in sample buffer as previously described [1 1], proteins were separated by 10% SDS/PAGE, and then transferred 97 to Immobilon transfer membrane (Millipore Corp., Bedford, MA, USA). Immunoblot analysis was carried out as described [1 1]. Briefly, the blots were blocked by incubating for 1 h at room temperature in Tris-buffered saline, pH 7.6, (20 mM Tris-HCL 137 mM NaCl) containing 0.1% Tween 20 (v/v) and 5% nonfat dry milk (w/v). To test for MAPK, the membranes were incubated for 1 h with the indicated anti-MAPK antibodies in Tris-buffered saline containing 0.1% Tween 20 and 5% nonfat dry milk, washed several times in this Tris- buffered saline, and then incubated with goat anti-mouse IgG as desired. Enhanced chemiluminescence (Amersham, Arlington Heights, IL) was used as a detection system according to the manufacturer's instructions. Microtubule-associated protein 2 (MAP-2) phosphorylation assay (MAPK activity assay) MAP-2 was purified from pig brain essentially as described [24]. The activity was assayed as described [11] but using MAP-2 as substrate instead of myelin basic protein . Briefly, cell lysates containing 2.5 pg proteins were incubated with MAP-2 (0.5 mg/ml), and [7-32P]ATP for 10 min at room temperature, the reaction was stopped by the addition of 5X sample buffer (1x sample bufier contains 0.05 M Tris-HCL pH6.9, 0.72 M mercaptoethanol, 9% glycerol, 2.3% SDS and 0.1% Bromphenol blue). The proteins in the reaction were separated on a 7.5% SDS-PAGE, and gels were dried and exposed to Kodak X-Omat film to visualize the amount of MAP-2 phosphorylation. Intracellular Ca2+ measurement The intracellular Ca2+ concentration was measured as described [41] both in the 98 resting cells and cells incubated with extracellular Ca2+ , lys-bradykinin, ionomycin, thapsigargin or EGF. Briefly, cells were cultured in Lab-Tek coverslip chambers and incubated in serum free medium (with 0.1 Ca”) for 24 hours. The cells were then loaded with a Ca2+ sensitive fluorescent dye, indo-l-AM (2 uM) in the same medium by incubating them for 1 hour at 37°C. To remove the extracellular dye, cells were then washed three times with McM medium (0.1 mM Ca2+) modified to contain 20 mM Hepes but without phenol red. They were then incubated in this medium until the beginning of the measurement (within 2 hours). The concentration of free Ca2+ in the cells loaded with indo-l-AM was calculated from data obtained with an ACAS 570 Interactive Laser Cytometer (Meridian Instruments, Okemos, MI, USA). An argon ion laser beam was led to the specimen on an inverted microscope, and emission from the illuminated spot on the sample was directed to a sensitive photomultiplier tube. The image pairs were captured at 8-second intervals. All image analysis experiments were conducted at room temperature. The ratio of the intensities of fluorescent emission at 405 nm and 530 nm (since the UV/Blue cube was used) with excitation at 355 nm was measured. The free Ca2+ concentrations in each single cell was calculated by comparing the ratio of the emissions to that generated in a standard curve. The standard curve was produced by mixing the indo-l , cell impermeate, with varying amounts of added Ca“ in a physiological buffer containing EGTA as described [42]. Cytosolic 1P3 measurement Cytosolic 1P3 was measured by using 1P3 binding protein [43] from 1P3 Biotrak Radioimmunoassay System Kit (TRK 1000) (Amersham Life Science Inc.). Samples were prepared essentially as described in the manufacturer’s instruction booklet with some 99 modifications. Briefly, cells were grown to confluence in 60 mm dishes in complete medium, and then incubated in serum free medium for overnight, treated with 10 mM Ca2+ for 1 min, or with 50 ng/ml bradykinin for 0.5 or 1 min, or left untreated. At the end of treatment, medium was aspirated and 0.5 ml of 10% cold trichloroacetic acid (v/v) was added to each dish, and samples in each dish were scraped by rubber policeman and transferred to an eppendorf tube. The precipitates were sedimented by centrifugation at 2100xg in microcentrifuge for 15 min at 4° C. The supernatant containing 1P3 was transferred to a 15 ml falcon blue cap tube and extracted three times with 10 volumes of water-saturated diethyl ether. 1P3 extracts were neutralized by titration to pH7.0-8.0 using NaHCO3. 1P3 concentrations were measured and calculated from the parallel standard curve made according to the manufacturer’s instructions. 100 RESULTS AND DISCUSSION Evidence that activation of MAPK by extracellular Ca2+ is partially inhibited by PMA Many protein growth factors stimulate the activation of PKC [24,44], which leads to the activation of downstream kinases, including MAPK [31]. Since in mouse fibroblasts, extracellular Ca2+ induces c-fos expression in a PKC-dependent manner [15], we first tested the involvement of PKC in the Ca2+ -induced activation of MAPK in human fibroblasts. Exposure human fibroblasts, as well as other types of cells, to PMA for a few minutes has been shown to activate PKC [45], and a long exposure to PMA downregulates PKC [20,24,45,46]. To determine whether PKC mediates the Ca2+-induced MAPK activation pathway in human fibroblasts, confluent LG] cells were incubated in medium overnight with or without 100 nM PMA , then treated for 5 min with Ca”, EGF, or fresh PMA, or left untreated. Cell lysates were assayed for MAPK activation. Incubation of cells with PMA for 5 min induced MAPK activation, as shown by the mobility shift of the p42 MAPK(Figure 1) and detected by an activity assay using MAP-2 as the substrate for phosphorylation by MAPK (data not shown). These data indicated that in human fibroblasts, MAPK activation is mediated by PKC. Down-regulation of the PMA-sensitive PKC by overnight exposure of the cells to with PMA completely abolished the effect of subsequent PMA treatment (Figure 1), indicating that PKC is completely downregulated by PMA . Overnight treatment of cells with PMA partially abolished Ca2+ -induced MAPK activation, but had no effect on EGF-induced MAPK activation, as shown by the p42 MAPK mobility shift (Figure 1) and detected by MAP-2 phosphorylation assay (data not shown). These results indicated that in human fibroblasts extracellular Ca2+-induced MAPK activation is 101 8:08 :09: 0 00:5 :0 0323:0380: 0:0 850:: 3:50: 0:... $0038: .80—00:08 Max—2-0:: :33 mix—0:0 :0_:0::88_ 0: 80.33 0:: 00:30:: 0:03 00:00.: :00 00808:: :2 :0 .38 m :0: ham 8%: cc: :0 .300 28 o— .<_2n_ 2: oo. :33 00:00:03 0:03 m:00 8:0: omém :83: .32: so: <2: 3053 :0 :33 0:: 30360.00 may 83008 00:: 830.: 0: 00w:0:0 83 83008 2:. .mUm $0: :33 83008 m.0_w0m 3 00:25:00 0: :30:w 0:03 2.00 .52.: E 0832.: 23:2: a: 3.6 52:38: E 0882 :c 5:303 : 2:0: 88888 8.8 0.8.28: _ :00 .80—<21- :00; + .. + I + .. u<_>_n_.::.m:n_ 102 partially dependent on PKC activation, whereas, EGF-induced MAPK activation is independent of PKC activation. Since PKC is completely downregulated by the overnight PMA treatment, these data suggest that there are at least two independent pathways for Ca” -induced MAPK activation, one dependent on a PMA-sensitive PKC, and one independent of this PKC. Evidence that extracellular Ca2+ does not induce c-raf-l phosphorylation Growth factors such as PDGF activate downstream kinases, such as PKC, raf-l , and MAPK, by phosphorylation [30,44,47]. It has been shown that the activation of PKC mediates c-raf-l kinase activation by direct phosphorylation [29,30]. The activated form can be detected after the separation of cellular proteins by SDS-PAGE since the phosphorylated form migrates more slowly [30]. To determine whether c-raf-l is in the pathway of extracellular Ca2+ -induced MAPK activation, we determined the phosphorylation status of c-raf-l in human fibroblasts before and after stimulation with Ca”, EGF, or PDGF. A slower moving form of c-raf-1(Figure 2) was observed in lysates from cells treated with PDGF, when compared with that from untreated cells (Figure 2), indicating that in human fibroblasts phosphorylation of c-raf-l is induced by PDGF. No mobility shift of c-raf-l was observed in lysates from cells treated with extracellular Ca2+ or EGF, indicating that extracellular Ca2+ does not act via a pathway involving c-raf-l. This conclusion agrees with that of Epstein [15], who showed that in mouse 3T3 cells, c-raf-l is not activated by extracellular Ca2+ but is activated by PDGF. 103 4:808:88 008:0::0: 3800:8003 00:5:0 038800080: 0:: :38: 38:0: 0:; 4:30:08 05 0:880: 0: ”:01??me fans : 830:0: x0038: 7:30.25: 8:: £828: :030::88_ 0: 803:: 0:: 00:80:: 0:03 8:33 :00 02:08:: ::0_ :0 .38 c: :0: mad—00: 8%: 310.38 m :0: “:00 8%: co: :0 a:0 28 o: :33 0080:: :0:: 0:: 8:0: 3 :0: 83008 00:: 82:0: 3 008803 0:03 803::00 0: :30:w £00 83818:: :0...— _.::.:-0 00:0 3 :0: 0000 $0 88:00:55— " 0::—ME Q. 3 :.ALN._ «Xavxflvmmo \o l04 Evidence that extracellular Ca2+ induces intracellular Ca2+ mobilization in human fibroblasts Intracellular Ca2+ has been shown to be an important second messenger in cell proliferation driven by protein growth factors and growth hormones [48-50]. Gill and colleagues [49,50] have shown that intracellular Ca2+ pools are directly linked to cell growth in hamster smooth muscle cells since when these investigators depleted the intracellular Ca2+ pools, the cells were arrested in G0 phase. To determine whether intracellular Ca2+ is mobilized as a result of incubating human fibroblasts with extracellular Ca”, the intracellular Ca2+ concentration was measured in resting cells and cells incubated with Ca“, lys- bradykinin, ionomycin, or EGF, using indo-l microflurometry with an ACAS fluorescent microscopy. Ionomycin, a calcium ionophore, and bradykinin, a protein mitogen which mediates its signal by binding to a receptor, were used as positive controls since it has been shown that the treatment of cells with these agents causes a transient increase in the free intracellular Ca2+ concentration [21] . As shown in Figure 3, cells treated with bradykinin (B) and ionomycin (C) exhibited a transient increase in the concentration of intracellular Ca2+, which lasted about 60 seconds. Afier the addition of Ca2+, intracellular Ca2+ concentration was increased quickly (Figure 3, D) to a level similar to that caused by ionomycin and bradykinin, and the transient lasted for about 180 seconds. No obvious intracellular Ca2+ transient was observed after EGF stimulation (Figure 3, E). This result is in agreement with a previous study in human fibroblasts that showed that when cells were pretreated with EGTA to chelate the extracellular Ca2+ in the medium [51] there was no noticeable elevation of the concentration of intracellular Ca2+ upon EGF stimulation. 105 U U g 2 A g 23 a 3 3 : r‘ 7‘ ~ g 1 g 1— I» oITlllIlTllTllllIlllll o jfi—rlllllllllljjlllj—l 0 100 200 300 400 O 100 200 300 400 3 1.0— 0.80 F "‘ 2 B A 3 E 0.60 :2" 3." § 1 ‘3‘ 0.4-1 0.2-4 olllllllllllllllllllll 0.0 lllrllrllllllllllll] O 100 200 300 400 O 100 200 300 400 30 1.0— : 0.8« s 25 c A G 3 - E 0.6- r' I r' "Q ~ "I 0.4: 2. 1— 9. q 02 «W 1 o TllllerlllllllllTll 0.0 IIITITlTlIllTrIIUIII O 100 200 300 400 0 100 200 300 400 3— 3 - 23 A 2 H 5. ~ ° 3 “' - 3" 9. 1'4 9.1 o TllllllllIIIIIIWTll] olllllTlTlllllllIlllTI O 100 200 300 400 O 100 200 300 400 Time (seconds) Time (seconds) Figure 3 Intracellular Ca2+ mobilization after Ca2+ stimulation. Subconfluent cells incubated in serum free medium overnight were preloaded with indo-l-Am for 1 hour, washed and placed in McM medium containing 0.1 mM Ca2+ without phenol red but with Hepes. lndo-l microflurometry were carried out usingACAS fluorescent microscope as described in “Material and method”. Intracellular Ca2+ concentrations were measured immediately following the addition of 0.1 mM Ca“ medium as control (A), 50 ng/ml lys-bradykinin (B), 1.5 ug/ml ionomycin (C), 10 mM Ca“(D), 100 ng/ml EGF (E), or 1 uyml thapsigargin (F), or measured after cells were pretreated with l ug/ml thapsigargin for 24-30 hours, then treated with l ug/ml thapsigargin (G), or 10 mM Ca’*(H ). Bach curve is the data taken from a single cell of3-5 cells measured in each experiment, and each experiment was repeated two tothree times with similar results. 106 Evidence that the mobilization of intracellular Ca2+ transient by extracellular Ca2+ is partially abolished by thapsigargin Because Gill and colleagues [50] have found that in mouse muscle cells, thapsigargin depletes the intracellular Ca2+ stores (pools) by specific inhibition of the endoplasmic reticulum (ER) Ca2+ -ATPase, and that this results in the arrest of the cells in G0 phase [49,50], we examined the effect of thapsigargin treatment on Ca2+-induced intracellular Ca2+ mobilization in human fibroblasts. We first measured the intracellular Ca2+ concentration in cells after 1 ug/ml thapsigargin treatment. To test whether prolonged treatment with thapsigargin depletes intracellular Ca2+ store, cells were incubated in medium with thapsigargin for 24-30 hours, and intracellular Ca2+ concentration was measured after the addition of thapsigargin. The addition of thapsigargin to human fibroblasts increased the intracellular Ca2+ concentration (Figure 3, F). The profile we obtained was similar to that showed by Chao et al. using human fibroblasts [21]. Prolonged treatment with thapsigargin totally blocked the thapsigargin-increased intracellular Ca2+ concentration (Figure 3, G). These data indicated that at the concentration used for the pretreatment, 1 ug/ml, the thapsigargin-sensitive Ca2+ pools are completely depleted. We then tested whether Ca2 + fiom the thapsigargin-sensitive stores is responsible for the intracellular Ca2+ increase after stimulation with extracellular Ca”. Cells were pretreated with thapsigargin for 24-30 hours, and intracellular Ca2 + concentration was measured afier the addition of extracellular Ca2 i. We found that thapsigargin pretreatment partially abolished the transient increase of intracellular Ca2 * concentration (Figure 3, H) that was normally seen after stimulation with extracellular Ca2+ (Figure 3, D) This indicated that the thapsigargin-sensitive Ca2 i stores are partially responsible for the Ca2 + increase after Ca2 * 107 stimulation. Previous reports suggest that the depletion of intracellular Ca2+ pools triggers Ca2+ entry across the plasma membrane by a process known as “capacitative Ca2+ entry” [3 2,52]. It is the Ca2+ from both the intracellular Ca2+ stores and from the Ca2+ entry channel [32] that accounts for the longer lasting increase in intracellular Ca2+ concentration after stimulation of the cells by various growth factors including PDGF [32,48]. The fact that we observed a longer lasting transient increase in intracellular Ca2+ concentration afier extracellular Ca2+ stimulation (Figure 3, D) than we saw in cells treated with ionomycin (Figure 3 C) or bradykinin (Figure 3, B) in low Ca2+ medium, and that this Ca2+ increase can be abolished only partially by thapsigargin pretreatment (Figure 3, H), suggests that, after extracellular Ca2+ stimulation, in addition to that Ca2+ from thapsigargin-sensitive intracellular Ca2+ stores, Ca2+ either from channels or from thapsigargin-insensitive stores, contributes to the transient intracellular Ca2+ mobilization we observed. It is known that there is more than one type of intracellular Ca2+ store in the endoplasm reticulum membrane [53,54] and that the thapsigargin-sensitive stores account for only 75% of the stored Ca2+ [53]. Evidence that activation of MAPK by extracellular Ca2+ is partially inhibited by thapsigargin To test whether in human fibroblasts MAPK activation is mediated by intracellular Ca2+ from the Ca2+ stores, we tested the effect of thapsigargin treatment on Ca2 +-induced MAPK activation. Cells were treated with thapsigargin for 24-30 hours, and then stimulated with Ca“, or with EGF, or were left untreated. Cell lysates were assayed for MAPK activation by immunoblot analysis. We found that thapsigargin partially inhibited the p42 108 .mEoEEaxo 3838 8:: E 3:830 0.63 3.32 3:88 $953.8 vE<$T==a 5:.» $9228 83956:: 2 80.33 203 mean». :00 60895:: c2 8 .58 n he mOm ER: 8— .5 Luv 28 o. 5:: e383 :2: 295 82:33.3 8m some E £3 2:. .32: 8: 52m 98 2.5»: ::_w8w_m%5 53 we 3:805 of .8 A2: 8: <35 he 28%: C Ewcmwfiawfi :8 8:035 .o 85QO of 5. fine: omém c8 8:62: out 833 E 32385 203 252.28 8 Siam 2.00 .5395: ¥m<2 v3.65...— Um .8 5.5 .3 385%.: 593332: ecu <5; .8 .59 «332: he 38.—o 2:. 9 239m i...a§ ’31:". Auxaszqu lllll++++l+Nm0.m®._ +|||..|Iu_mum.H ._. OFEoE +uuu++uun<§a +++ | ++ l l + l + l I 109 MAPK mobility shifi induced by extracellular Ca2+ , but did not inhibit the mobility shift induced by EGF (Figure 4). These data indicate that intracellular Ca2+ mobilization from the thapsigargin-sensitive Ca2+ stores lies, at least in part, in the pathway between extracellular Ca2+ and MAPK activation, but not in the pathway between EGF and MAPK activation. To determine whether the PMA-sensitive PKC and intracellular Ca2+ from the ' thapsigargin-sensitive stores are in the same or a different pathway for extracellular Ca2+ - induced MAPK activation, cells were pretreated with PMA and thapsigargin together and then stimulated with Ca2+ for 5 min and lysates were subjected to immunoblot analysis using anti p42 MAPK antibody. The pretreatment with PMA and thapsigargin resulted in an almost complete inhibition of Ca2+-induced MAPK mobility shift (Figure 4), the inhibition was greater than by either of the inhibitors given alone. This indicates that PMA-sensitive PKC and thapsigargin-sensitive intracellular Ca2+ stores lie, at least partially, in two distinct pathways. EGF-induced MAPK activation is not affected by either PMA or thapsigargin, indicating that EGF does not act through PKC activation or mobilization of the thapsigargin- sensitive intracellular Ca2+ stores. Evidence that activation of MAPK by extracellular Ca“ is abolished by the channel blocker lanthanum, but not by the channel blockers verapamil or nifedipine Growth factors such as PDGF evoke transient increase in intracellular Ca2+ concentration through mobilization of intracellular Ca2 + stores and Ca2 + entry by channels [48]. Ca” channel antagonists have been shown to decrease the rate of growth of many cell types [48,55]. Estacion and Mordan [48] showed that in mouse fibroblasts the selective inhibition of the sustained increase in Ca2+ concentration b blockin Ca2+ influx with Y g 110 .mEoEEuaxu EchotG 0.5 E UoESno 20>» $.32 3:E_w Scoot—8 ¥m “53:56 8 Esau mzou £22333 ¥m<2 rouse E-..— Um .3 :«U .3 «mi—cunt; ._.—238.: ._o «outm— 9 9:33 1" .‘g‘txagmva ||++uu+mm0 ++----n_om + .. + I + I 5.3591 4.8:. ”4 to Ca” or EGF. We found that, like W-7, calmidazolium (15 uM) partially abolished both Ca2+ and EGF-induced MAPK shifi (data not shown), whereas W-12 (9O uM), the inactive analogue of W—7, had no obvious effect on Ca2*- or EGF-induced MAPK activation (data not shown). Since two different calmodulin antagonists show a similar effect, and W-12, the inactive analog of W—7, which has approximately 10 times lower affinity for calmodulin than W-7 [63,64,66,67], shows no obvious effect on MAPK activation, we conclude with Takuwa et al. [63,64] that the effect of calmodulin antagonists is very likely based on their antagonism against calmodulin. Lu and Means conclude in a review [68] that Ca2+ is an absolute requirement for all enzyme activating functions of vertebrate calmodulin. This supports our suggestion that calmodulin mediates the function of intracellular Ca2+ afier extracellular Ca2+ stimulation. We have no ready explanation for the fact that EGF-induced MAPK activation is also abolished by calmodulin antagonists, although no obvious intracellular Ca2+ elevation was observed after EGF stimulation (Figure 3). It is also possible that calmodulin antagonists have effects on other components in the signal transduction pathway, although such a mechanism is not supported by W-12 data. The role of PLC In mammalian cells, protein growth factors activate G protein-coupled receptors or tyrosine kinase-linked receptors, which stimulate the production of DAG and 1P3 through the activation of PLC [69]. DAG activates PKC, whereas 1P3 increases cytosolic Ca2+ level by triggering the release of Ca2+ from intracellular stores. These two receptor-dependent mechanisms are coupled to different isoforms of the PLC family. G protein-coupled ”5 receptors activate PLCB [69-72]. The tyrosine kinase-linked receptors act through PLCyl by tyrosine phosphorylation [70,73]. In human and mouse keratinocytes and bovine parathyroid cells, extracellular Ca2+ has been shown to activate PLC and produce DAG and IP3, which in turn, activate PKC and elevate intracellular Ca2+ levels [2,3]. The fact that our results show that both PKC activation and intracellular Ca2+ elevation from intracellular Ca2+ stores are involved in the extracelluar Ca2+-induced MAPK activation pathway, suggests that PLC is involved in the activation of PKC and elevation of intracellular Ca2+. To test this hypothesis, we measured PLC activity by the accumulation of IP3 in cells [74], using an IP3 binding protein [33]. Figure 7 shows that IP-3 level was elevated 8-10-fold after 1 min of Ca2+ stimulation, and 5-fold after bradykinin stimulation for 30 seconds, and 3-fold after 1 min stimulation. The increased level found after bradykinin treatment is similar to that shown in other studies with human fibroblasts [75,76]. Bradykinin is known to function through the G protein-coupled bradykinin receptor and activate PLCB to generate IP3 and DAG [72,77]. The elevation of IP3 afier Ca2+ stimulation suggests that PLC is involved. Since PLCy has been shown to be phosphorylated on its tyrosine residue upon activation [70,73], we tested its involvement by immunoprecipitation followed by immunoblot analysis. Human fibroblast LGl cells were treated with Ca2+ (Figure 8, lane 2), or EGF (Figure 8, lane 3), or lefi untreated (Figure 8, lane 1), and the cell lysates were immunoprecipitated with anti-PLC'y antibody as described [78]. Proteins were subjected to SDS-PAGE, transferred to a membrane, and probed with anti-phosphotyrosine antibody for immunoblot analysis. Human epidermoid carcinoma A431 cells which are known to exhibit PLCy activity when treated with EGF [78] were used as a positive control. Our results show that EGF induced PLCy tyrosine phosphorylation in A431 cells (Figure 8, lane 5). No ll6 N lllllllJlllllL A IP3 (pmole) Figure 7 Cytosolic level of IP3 was increased after Ca” stimulation. IP3 was measured as described in “Material and methods” in cells with no treatment (1), or cells treated with 10 mM Ca2+ for l min (2), or 50 ng/ml bradykinin for 0.5 min (3) or 1 min (4). Values are means :hS.E.M. of four observations. ll7 .858..on 0 822.3 03. :. .3223 203 9.3.2 2.E.m avenge 0:.m2bofimosa-..:m 2...: £2.35. 830:3:8. o. c.3833 :2: .3 ”:2. 33.2: 65.02.20: .0 Am. 8:2. 33.2: >01... .2: 5.3 uo.a..n.u8ao:=EE. :5 622.2: 29$ 8.3... :60 Av .. 8:2. 3.85:: co. 8 .:.E m .o. 3 .m .m 8:2. .0m 8 AN 6:2. $5 5.3 :83: :2: “En £52. em 8.. E335 no... E33 :. 3.2.3:. 263 8:03.:8 o. :32w 3-.» 852.218 .mv< 8 Am -. 3:2. 2.8 .O‘. 62.2.32252... >0...— :o .Nau 3.3.3.2.: .e 39...”. a «._.—ME 0.0.. mm r, vmv< v0.— 118 tyrosine phosphorylation of this protein was observed upon Ca2+ (Figure 8, lane 2) or EGF (Figure 8, lane 3) stimulation in human fibroblasts. This indicates that PLCy is not involved in Ca2+ -induced MAPK activation pathway and that some other PLC isoform, such as PLCB, is involved. Evidence that MAPK activation by extracellular Ca“ is partially inhibited by pertussis toxin Because of the increase of cytosolic level of IP3 afler Ca2+stimulation, which suggests the involvement of PLC, we hypothesized that PLC might be activated by a G protein. Based on their sensitivity to pertussis toxin, G proteins are divided into two categories [70]: pertussis toxin-sensitive and -insensitive G-protein. Therefore, we tested the effect of pretreatment with pertussis toxin on Ca2+-induced MAPK activation. Cells were pretreated with 0.5 ug/ml pertussis toxin overnight, and then stimulated with Ca2+ or EGF, and their lysates were assayed for MAPK activation. Pertussis toxin partially abolished the MAPK mobility shift caused by extracellular Ca“, but had no effect upon EGF stimulation (Figure 9). This suggests that Ca2+ -induced MAPK activation is at least partially dependent on a pertussis toxin sensitive G-protein, whereas EGF-induced MAPK activation is independent of this type of G-protein. These data indicating the involvement of a pertussis toxin sensitive G protein, which has been shown to be one of the major activators of PLC, and the data provided above indicating the involvement of PKC, which can be activated by DAG, along with the IP3 elevation after Ca2+ stimulation, strongly suggest that PLC is involved. 119 .m.:6E..oqxo 26.656665 66.5 :. 6626.66 6.63 3.666.. ..2..:.m $665.5. v...<.>.-..:6 5.3 29:65. 3.66.388. 6. .6633. 6.63 66.3... :60 66.66.56 ...6. ..6 .:.E m .6. "am. ..6 $60 6. 6666...... :65 6:6 .366... em .6. GEE: n... :68. 232.2. 365.3 .6 5.3 86.66.: 66... £6.66. :. 66.6666... 263 .:6.._..:6o 6. :BEw 2.60 .636. 6.33.2. E 633...... 2.2:... a. :5 52.32:; 2 6.3.: 2 5.6.3.8.. 2; a as»... ‘ ‘u ‘ - AI¥Q<§NVQ I I + + |+Nm0 + + - - -uom. + n + u ..x.E .591 .66... l20 Evidence that the Ca2+ sensing receptor of Brown et al. [33] is not involved in extracelluar Ca2+-induced cell growth or MAPK activation in mouse skin fibroblasts How the fibroblasts sense the extracellular Ca“ concentration is unknown. It could be that Ca2+ channels open when the extracellular Ca2+ reaches a critical threshold, as suggested by the inhibition effect of lanthanum. On the other hand, the involvement of a G protein suggests that there may be Ca2+ specific receptors on the external cell surface, analogous to protein growth factor receptors, and that these are involved in this process. Just such a Ca2+-sensing receptor has been cloned from bovine parathyroid cells and has been shown to mediate the elevation of intracellular Ca2+ and IP3 [33]. Mice lacking this receptor show a dramatically retarded growth [35], which suggests that this receptor may play a role in growth control. Therefore, we tested the possible involvement of this receptor in regulating cell growth and MAPK activation by using knockout mice lacking this Ca2+ sensing receptor [3 5]. After determining that mouse skin fibroblasts exhibit the same ability as human skin fibroblasts to grow in serum free medium with 1 mM Ca2+ and to induce MAPK activation after Ca2+ stimulation (data not shown), we tested whether skin fibroblasts from mice lacking this receptor exhibit any deficiency in growth or MAPK activation after Ca2+ stimulation, when compared with their wild type or heterozygous counterparts. We found that all mouse skin fibroblasts, regardless of their receptor status (+/+, +/-, or -/-), had the same ability to grow in the medium containing 1 mM Ca2+ or supplemented with 10% FBS (data not shown). Also 10 mM extracellular Ca2+ or 10%FBS induced a MAPK mobility shift in these mouse fibroblasts to the same degree, regardless of the status of the Ca2+ sensing receptor (Figure 10). These results indicate that this Ca2+ sensing receptor is not involved in Ca2+-induced cell growth or MAPK activation in mouse fibroblasts. Since l2l 5:68.598 63. :. 568636 6.63 3.68. 8:85 $666.8: Max:258: 5.3 £9268. 56.66::88. 6. 86.33. 263 356m». :00 5235:: an. :6 .99.. $0. :6 Kano 28 o. 8 63698 :65 6:: .Ew.:._o>6 83.668 66.... 83.—om :. 58336:. 263 36066: .266. .6366»: 9:83 :60 65 :6. -\- .-\+ .+\+ 263 55.3 $.06 5:625:60 ...m66568 6:6 8.58 E: :. wontomow 3 6685.3 6:: .6238: 263 3minofic :.6_m 3:62 3836.5: 86: 62.68 :. mm..— ..6 :60 88:39:: .3 :625366 v-mSZ o— 9...»...— AIXQ