.7. £31.: an 1“? Exam“... 5.: 3. Ii: 4.333%”, W fiqmfiflflffl‘lwn.i : . Abra; finflififi 1157.3}fi. r3...» . ls . u. dun"; H. “w“: . ‘12:. J... i . p ‘3‘: shin}. It'lljlu 1001/ LiBRARY Michigan State University This is to certify that the dissertation entitled Molecular Mechanism of Fumonisin-induced Kidney Toxicity \ presented by Min Sun Kim has been accepted towards fulfillment of the requirements for Ph.D Human Nutrition-Environmental degree in Toxicology I Major professor Date [/Ll/OZ. MS U is an Affirmative Ardour/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE ' DATE DUE 6/01 c:/ClRC/DateDue.p65.p.15 MOLECULAR MECHANISM OF FUMONISIN-INDUCED KIDNEY TOXICITY By Min Sun Kim A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition and Institute for Environmental Toxicology 2002 ABSTRACT MOLECULAR MECHANISM OF FUMONISIN-INDUCED KIDNEY TOXICITY By Min Sun Kim F umonisin B, is a toxic and carcinogenic mycotoxin with renal tubular epithelial cells being the most sensitive target. Due to structural similarities with sphingoid bases, fumonisin B, inhibits ceramide synthase in the de novo biosynthetic pathway of sphingolipids causing depletion of complex sphingolipids and accumulation of sphingoid bases. The goals of this research were to use LLC-PK, renal tubular epithelial cells to: 1) Determine whether fumonisin B, kills cells by inducing apoptosis and evaluate the role of disruption of sphingolipid metabolism in fumonisin-induced cell death; 2) Identify genes affected by fumonisin B, and investigate the molecular mechanism whereby fumonisin B, induces gene expression; and 3) Determine whether fumonisin B, affects signaling pathways involved in cell proliferation and cell death. Fumonisin B, produced morphological changes and time-dependent increases in fragmented DNA indicative of apoptosis within 16 hours and simultaneously caused accumulation of sphinganine. To investigate the role of sphinganine in fumonisin B,- induced apoptosis, B-fluoroalanine (BFA) was used to inhibit serine palmitoyltransferase, which catalyzes an earlier step in the sphingolipid biosynthetic pathway. BFA blocked sphinganine accumulation and prevented fumonisin B,-induced DNA fragmentation, confirming that fumonisin B,-induced apoptosis depends on accumulation of sphinganine. Fumonisin B, selectively induced several genes including calmodulin; whereas, the mycotoxin did not affect the bcl-2 family genes bcl-2, bcI-x, and bax. Fumonisin B, increased both calmodulin mRN A and protein in concentration-dependent manners, and the calmodulin antagonist W7 blocked fumonisin B,-induced DNA fragmentation, supporting a role for calmodulin in fumonisin B,-induced apoptosis. F umonisin B, did not influence stability of calmodulin mRNA and appears to induce calmodulin by increasing transcription. BFA added together with fumonisin B, blocked sphinganine accumulation and blunted induction of calmodulin, suggesting a role for sphinganine or the l-phosphate metabolite. Exogenous addition of sphinganine had no effect on calmodulin expression, but addition of the l-phosphate significantly increased calmodulin mRNA, indicating sphinganine 1- phosphate mediates fumonisin B, induction of calmodulin. To examine signaling pathways involved in fumonisin-induced cell death, LLC-PK, cells were cultured with either fumonisin B, or sphinganine. Exogenously added D-erythro- sphinganine acted more rapidly than fumonisin B,, inhibiting growth and causing cell death. Both fumonisin B, and sphinganine reduced phosphorylation of MEK and ERK without affecting phosphorylation of Raf. Fumonisin B, and sphinganine also decreased phosphorylation of Akt (protein kinase B). Furthermore, addition of BFA together with fumonisin B, blocked fumonisin-induced dephosphorylation, suggesting that sphinganine mediates the effect of fumonisin B, on phosphorylation of MEK, ERK, and Akt proteins. Taken together, fumonisin B, appears to cause kidney toxicity by disrupting sphingolipid metabolism and causing accumulation of sphinganine and sphinganine 1- phosphate. These sphingolipid metabolites induce selected genes including calmodulin, and alter signaling through cell proliferation and cell death pathways leading to renal tubular epithelial cell apoptosis This dissertation is dedicated to my devoted parents for their love and encouragement. iv ACKNOWLEDGEMENTS First of all, I thank my advisor Dr. Joseph J. Schroeder, without whom I would not be able to get this far. He provided support, kindness, assistance and friendship in every step I make. I respect him as a researcher as well as a human being. I will remember him as my true mentor no matter where I am. I also would like to thank my guidance committee members, Drs. Dale R. Romsos, Maurice R. Bennink, and Patricia E. Ganey for their truly valuable advice, criticisms, and comments which have allowed me to improve as a scientist. I am also gratefiJl to Dr. Thomas Wang for his assistance in quantitative RT-PCR, and to Dr. Doh-Yeel Lee for his technical guidance for differential display RT-PCR. I thank member of the Food Microbiology research group for their enduring kindness in sharing their equipments and knowledge. I would like to thank both present and past lab members Eun-Hyun Ahn, Eun-Hee Lee, Hong Yang, Chi Zhang, and Jason Weisinger for their technical assistance, helpful comments and friendship. My special thanks go to my brothers and sister for their encouragement, patience, kindness and support for all these years. Without them, I wouldn’t be able to focus on my studies and research. To my parents, I can’t say enough to express my gratefulness and appreciation. I thank them for letting me be who I am and do what I want to do. TABLE OF CONTENTS LIST OF FIGURES ...viii ABBREVIATIONS ...x INTRODUCTION 1 CHAPTER I. LITERATURE REVIEW 3 A. FUMONISINS ...4 A. 1 . Introduction ...4 A2 Toxicity 5 A3. Carcinogenicity ...6 A4. Structure ...8 A5. Toxicokinetics 10 A6. Metabolism ...1 1 A7. Inhibition of Ceramide Synthase ...11 B. SPHINGOLIPIDS 14 B. 1 . Introduction 14 82. De nova Biosynthesis ...15 B. 3. Turnover ...17 B. 4. Sphingolipid Signaling ...17 B. 5. Disruption of Sphingolipid Metabolism as a Causative Factor in Fumonisin-induced Toxicity .26 C. LLC-PK, RENAL TUBULAR EPITHELIAL CELLS AS A MODEL ...30 D. HYPOTHESIS AND SPECIFIC AIMS ...31 CHAPTER II. F UMONISIN B, INDUCES APOPTOSIS IN LLC-PK, RENAL EPITHELIAL CELLS VIA A SPHINGANINE- AND CALMODULIN- DEPENDENT PATHWAY ...32 A. ABSTRACT ...33 B. INTRODUCTION ...34 C. MATERIALS AND METHODS ...36 C. 1. Reagents and cell culture ...36 C. 2. Analysis of DNA fragmentation by agarose gel electrophoresis" 3.6 C. 3. Flow cytometric analysis of apoptotic cells ...37 C4. Mass measurement of sphinganine ...37 C5. RNA isolation ...38 C. 6. Quantitation of bcl-2, bcI-x, and box mRNA levels ...38 C. 7. Densitometry analysis ...39 C. 8. Differential display reverse transcriptase polymerase chain reaction (DDRT-PCR) ...39 C9. Northern analysis ...39 C.lO.cDNA cloning ...39 vi C .1 1. DNA sequencing C.12. Immunoblot analysis for calmodulin C. l 3. Cell protein measurement C.l4. Statistical analyses D. RESULTS E. DISCUSSION ...40 ...40 ...40 ...40 ...4] ...53 CHAPTER III. SPHINGANINE l-PHOSPHATE MEDIATES CALMODULIN EXPRESSION INDUCED BY F UMONISIN B, IN LLC-PK, PORCINE RENAL TUBULAR EPITHELIAL CELLS A. ABSTRACT B. INTRODUCTION C. MATERIALS AND METHODS C.1. LLC-PK, cell culture C.2. Materials C.3. Analysis of DNA fragmentation by agarose gel electrophoresis. C.4. Differential display reverse transcriptase polymerase chain reaction (DDRT-PCR) C.5. Immunoblot analysis for calmodulin C.6. Northern blot analysis C.7. Statistical analyses D. RESULTS E. DISCUSSION CHAPTER IV. FUMONISIN B, AND SPHINGANINE REDUCE PHOSPHORYLATION OF MAPK/ERK AND PKB/AKT IN LLC-PK, PORCINE RENAL TUBULAR EPITHELIAL CELLS A. ABSTRACT B. INTRODUCTION C. MATERIALS AND METHODS C. l . Materials C.2. Cell culture C.3. Total nucleic acid measurement C.4. Western blot analysis C.5. Statistical analyses D. RESULTS E. DISCUSSION CHAPTER V. SUMMARY AND CONCLUSIONS APPENDICES LIST OF REFERENCES vii ...62 ...63 ...64 ...66 ...66 ..67 .67 ...67 ...68 ...68 ...69 ...69 ...79 ...85 ...86 ...87 ...91 ...9] ...91 ...91 ...92 ...92 ...92 ...98 ...105 ...l 10 ...116 Figure 1.1. Figure 1.2. Figure 1.3. Figure 1.4. Figure 1.5. Figure 1.6. Figure 1.7. Figure 1.8. Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 2.8. Figure 2.9. Figure 2.10. LIST OF FIGURES Structures of fumonisins and related mycotoxins Structures of fumonisin B, and sphinganine De nova sphingolipid biosynthesis Sphingolipid turnover Sphingolipid signaling and cellular effects Structures of sphingoid bases Effects of sphingolipids on cell proliferation signaling pathway Effects of sphingolipids on apoptotic signaling pathway Fumonisin B, induces death of LLC-PK, kidney cells Fumonisin B, causes sphinganine accumulation in LLC-PK, kidney cells B—Fluoroalanine blocks sphinganine accumulation and fumonisin B,-induced DNA fragmentation B—Fluoroalanine blocks fumonisin B,-induced DNA fragmentation Fumonisin B, does not affect gene expressions of bax, bc1-2, and bcl-x Differential display of cDNAs from kidney cells Fumonisin-B, induces expression of K3 mRNA Homology between K3 DNA sequences and human calmodulin B-Fluoroalanine reduces fumonisin-induced calmodulin expression F umonisin B, increases intracellular level of calmodulin protein in a dose-dependent manner viii ...12 ...16 ...18 ...20 ...22 ...24 ...29 ...42 ...43 ...45 ...48 ...49 ...50 ...51 ...52 ...54 Figure 2.11. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5. Figure 3.6. Figure 3.7. Figure 3.8. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 4.8. Figure 5.1 Calmodulin antagonist, W7, blocks fumonisin-induced apoptosis in LLC-PK, cells De nova biosynthetic pathway of sphingolipids and inhibition of ceramide synthase by fumonisin B, Fumonisin B, increases calmodulin mRNA and calmodulin protein in a concentration-dependent manners Fumonisin B, does not affect stability of calmodulin mRNA Novel gene transcription is required for fumonisin B,-induced apoptosis Fumonisin B, selectively induces calmodulin Depletion of complex sphingolipids does not affect expression of calmoduolin Sphingoid bases does not mediate fumonisin-induced calmodulin expression Sphinganine l-phosphate induces calmodulin expression Structures of fumonisin B,, sphinganine, and sphingosine Fumonisin B, inhibits growth and induces death of LLC-PK, kidney ...93 cells Sphinganine inhibits growth and induces death of LLC-PK, kidney cells Time-dependent changes in the phosphorylation of ERK in control and fumonisin B,-treated LLC-PK, cells Fumonisin B, reduces phosphorylated of ERK and MEK, but not Raf-1 Sphinganine reduces phosphorylated ERK, MEK, but not Raf-l F umonisin B, reduces phosphorylated Akt. Sphinganine reduces phosphorylated Akt Summary of possible molecular mechanism of fumonisin-induced kidney toxicity ix ...55 ...65 ...70 ...72 ...73 ...75 ...76 ...77 ...78 ...89 ...93 ...95 ...97 ...97 ...99 ...99 ...109 Abbreviations pg AGP BSA Ca CAPP CYP2C1 1 CytoC EDTA ERK F ADD FBS g G3PDH hrs IKB IL iv kg MAPK MEK mg min ml NFKB PCR PDGF PI3K PKB PKC pmol RT ser SK thr TNF ABBREVIATIONS Meaning Microgram 0t,-acid glycoprotein Bovine serum albumin Calcium Ceramide-activated protein phosphatase Cytochrome P450 2C 11 Cytochrome C Ethylenediamine tetraacetic acid Extracelluar receptor activated kinase Fas-associated death domain Fetal bovine serum Gram Glyceraldehyde 3-phosphate dehydrogenase Hours Inhibitory K B Interleukin Intravenous Kilogram . Mitogen-activated protein kinase MAPK kinase Milligram Minute Milliliter Nuclear factor K B Polymerase chain reaction Platelet-derived growth factor Phosphatidylinositol 3-kinase Protein kinase B Protein kinase C Picomole Reverse transcriptase Serine Sphingosine kinase Threonine Tumor necrosis factor INTRODUCTION Fumonisins are toxic and carcinogenic mycotoxins produced by F usarium moniliforme, the most common mold found on corn throughout the world (Marasas, 1996). Fumonisins cause animal diseases (Wilson et al. , 1990; Harrison et al. , 1990) and have been associated with esophageal (Marasas et al., 1988), liver (Ueno et al., 1997) and stomach (Groves et al., 1999) cancers in humans. The mechanism of action of fumonisins is not clear, but appears to involve disruption of sphingolipid metabolism. Due to the structural similarity with the sphingoid base backbones of sphingolipids, fumonisins inhibit ceramide synthase (sphingoid N- acyltransferase) which acylates sphinganine to form ceramide in the de novo biosynthetic pathway of sphingolipids. Inhibition by fumonisins results in accumulation of sphinganine and depletion of complex sphingolipids (Wang et al. , 1991 ; Merrill et al. , 1993a). Complex sphingolipids mediate cell-cell, cell-substratum, and cell-microbial interactions and serve as substrates for agonist-induced cell signaling; while, sphingolipid metabolites serve as bioactive molecules which affect cellular behavior by acting as second messengers for many growth factors and cytokines. Sphingoid bases, which include sphinganine, have been shown to arrest cell growth and induce apoptosis, early responses in tissues exposed to fumonisins (Voss et al. , 2001 ). Apoptosis may mediate fumonisin-induced toxicities in liver and kidney, the latter being the most sensitive organ (Bondy et al., 1996; Bucci et al., 1998). Sphingoid bases modulate expression of bcl-2 family genes (Sakakura et al., 1996; Shirahama et al. , 1997) and decrease activity of MAPK (Raeder et al. , 1999) to suppress cell proliferation and induce apoptosis, suggesting that these pathways may be affected by fumonisins. In addition, the PI3K/Akt pathway is a probable target of fumonisins because ceramide (Schubert et al., 2000) and sphingosine (Chang et al., 2001) inactivate Akt to induce apoptosis. In the present study, LLC-PK, porcine renal tubular epithelial cells were used as a model system to examine the type of cell death, the role of disruption of sphingolipid metabolism and effects on gene expression and cell signaling in the molecular mechanism of fumonisin-induced kidney toxicity. CHAPTER 1 LITERATURE REVIEW A. Fumonisins A. 1. Introduction First discovered in 1988 (Gelderblom et al., 1988), the fumonisins are a family of cytotoxic and carcinogenic mycotoxins produced by F usarium moniliforme and other related fungi which are common contaminants of corn and other agricultural commodities occurring in the field and during storage in the United States and throughout the world (Marasas, 1996). More than ten types of fumonisins have been isolated and characterized with the ‘B’ series representing the most common of these mycotoxins. Fumonisin B, is the most prevalent form and has been shown to cause all of the animal diseases associated with consumption of F. moniliforme culture materials (Riley et al., 1993b); whereas, fumonisin B2 which is less prevalent exhibited a higher toxicity and more specific binding to rat hepatocytes than fumonisin B, (Cawood et al., 1994). The primary concern about fumonisins is due to the prevalence of Fusarium moniliforme as a common contaminant of corn. Naturally occurring fumonisins have been documented in corn (Sydenham et al., 1990) and com-based animal feeds (Plattner et al., 1991). The extent of contamination of raw corn with fumonisins varies with geographic location, agronomic and storage practices, and the vulnerability of the plants to fungal invasion during all phases of growth, storage, and processing. High concentrations of fumonisins are associated with hot and dry weather followed by periods of high humidity (Shelby et a1 ., 1994). Numerous studies have reported that fumonisins naturally contaminate corn at concentrations ranging from <1 to 330 [Lg/g for fumonisin B, and concentrations ranging from <1 to 48 pg/g for fumonisin B2 which were associated with outbreaks of porcine pulmonary edema in the United States from 1989 to 1990 (Colvin and Harrison, 1992; Haschek et al., 1992; Osweiler et al., 1992; Ross et al., 1992; Ross et al., 1991). Furthermore, recent analyses of food-grade corn and com-based food products of United States origin for human consumption showed that 71% of the samples contained fumonisin B,, ranging from 43 to 1642 pg/kg (Gutema er al., 2000). The results from this report suggest that United States consumers are at risk for fumonisin-induced health problems. A. 2. Toxicity Fumonisins are associated with a variety of adverse health effects in livestock animals. Incidence of leukoencephalomalacia in horses has been linked with mold- contaminated feeds (Wilson et al., 1990) and was reproduced by feeding a horse naturally contaminated corn screenings (Marasas et al., 1989a) and by orally dosing fumonisin B, (Kellennan et al., 1990). Fumonisin B, also causes pulmonary edema in pigs (Harrison et al., 1990; Osweiler et al., 1992). Pulmonary edema which was preceded by respiratory distress was induced in pigs fed corn screenings containing 155 mg/kg fumonisin B, (Harrison et al., 1990; Motelin et al., 1994) and in one pig by intravenous injection of 11.3 mg fumonisin B, (Colvin and Harrison, 1992). In addition to lung, both liver and kidney are target organs of fiJmonisin B, in swine (Haschek et al., 1992). Fumonisins also are toxic for laboratory animals. Progressive toxic hepatitis characterized by hepatocellular necrosis, bile duct proliferation and fibrosis was observed in rats treated with fumonisin B, in short-term toxicity tests (Gelderblom etal., 1991). The chromatin of exposed cells is irregularly condensed and marginated or may be fragmented, a common feature of apoptotic cells. Also, necrotic hepatocytes are present hours after exposure to the mycotoxin. Serum chemical indications of hepatocellular injury, including increased alanine and aspartate transaminase, alkaline phosphatase, and lactate dehydrogenase activities as well as increased cholesterol and triglyceride concentrations are routine, early findings (Voss etal., 2001). In addition, fumonisin-induced kidney toxicity was reported in rats (Bondy et al., 1996), lambs (Edrington er al., 1995), and rabbits (Bucci et al., 1998). Whereas hepatotoxicity was observed with a diet containing 150 mg/kg, renal toxicity was observed even at concentrations as low as 15-50 mg/kg of fumonisin B, (Voss et al., 1993 ), suggesting that kidney is more sensitive to fumonisin-induced toxicity than liver (Bondy et al., 1996; Bucci et al., 1998; Voss et al. , 2001). In Sprague-Dawley and Fischer 344 rats fed fumonisin B,, males were more sensitive than females (Voss et al. , 2001). As in liver, apoptosis is the initial microscopic finding in the kidney and apoptotic cells are initially found exclusively in tubules of the outer medulla. Mitotic figures appear and the number of apoptotic cells increases in the tubule epithelium as tissue injury progresses (Voss etal., 2001). It should be noted that simultaneous cell loss and replacement has been observed including apoptosis and mitosis at the cellular level, tubular atrophy and hyperplasia on the histologic level, and grossly by decreased kidney weight (Voss et al., 2001). The replacement or regeneration process accompanying apoptosis may mediate, at least in part, fumonisin-induced carcinogenicity. A. 3. Carcinogenicity F umonisin B, is carcinogenic in animals. This mycotoxin causes primary hepatocellular carcinoma and cholangiocarcinoma in rats at a dietary concentration of 50 mg/kg fed for 18-26 months (Gelderblom et al., 1991). A recent carcinogenicity study by the National Toxicology Program demonstrated that fumonisin B, induces renal tubule tumors in male F344 rats and hepatic tumors in female B6C3F mice (Howard et al., 2001). Fumonisin B, may affect human health in certain areas of the world where corn is a staple of the diet. Epidemiological studies suggest that there is a strong correlationship between consumption of fumonisin B, and the incidence of human esophageal cancer in certain areas of South Africa (Marasas et al., 1988b) and liver and stomach cancers in China (Ueno et al., 1997; Groves et al., 1999). The high incidence of esophageal cancer in the Transkei region of South Africa was associated with the consumption of beer brewed with moldy corn and a nonalcoholic fermented drink made with corn (Rose, 1981). In a sample of home-grown corn from a high incidence area of esophageal cancer in the Transkei, visibly healthy kernels contained 44 pg/ g fumonisin B, and F usarium-infected kernels contained 83 [lg/g fumonisin B, (Marasas, 1996). Esophageal basal cell hyperplasia has been demonstrated in rats by feeding culture material of F. moniliforme (Marasas et al., 1984), providing a partial explanation for the esophageal cancer epidemic in certain regions of the world. The mechanism(s) by which fumonisins cause cancer is not clear. F umonisins were neither mutagenic in the Salmonella mutagenicity assay (Gelderblom et al., 1991) nor genotoxic in DNA repair assays using primary rat hepatocytes (Norred et al., 1992). F umonisin B, was found to be a weak cancer-initiating agent when rats were fed fumonisins at a level of 100 mg/kg diet for 21 days, for which the intact molecule and the presence of a free amino group was essential (Gelderblom et al., 1993). In addition, fumonisin B, was demonstrated to be a strong cancer-promoter for rat liver when diethylnitrosamine was used as a cancer initiator (Gelderblom et a1 ., 1996b). F umonisin B, exhibited a cancer-promoting effect in the absence of adverse hepatotoxicity and at dietary levels that did not show any sign of cancer-initiation. The carcinogenic effects of fumonisins may be due to their mitogenicity indicated by stimulated DNA synthesis as shown in Swiss 3T3 cells (Schroeder et al., 1994). Recent advances in the molecular genetics of cancer suggest many nongenotoxic carcinogens may increase the risk of various types of genetic errors by stimulating cell division per se (Preston-Martin et al., 1990). For example, cell division is necessary for conversion of adducts or other single stranded DNA damage to gaps or mutations (Ames and Gold, 1990; Cohen and Ellwein, 1990). In addition, cell division allows for mitotic recombinations that result in changes that are more profound than those from a single mutation (Ramal, 1988). Therefore, it is possible that fumonisins might promote tumor formation by stimulating cellular proliferation. A. 4. Structure F umonisins are categorized as members of either the ‘A’, ‘B’, or ‘C’ series. The ‘B’ series is the most abundant and toxic of the fumonisins. Fumonisin B, is the diester of propane-1,2,3-tricarboxylic acid and 2-amino-12,16-dimethyl-3,5,10,14,15- pentahydroxyicosane (Figure 1.1). The propanetricarboxylic acid moieties are esterified at the C-14 and C-15 positions of the 20-carbon arninopentol backbone. Less is known about fumonisins in the ‘A’ and ‘C’ series. Fumonisins in the ‘A’ series have an amino group which has been acetylated, and they appear to be relatively innocuous (Gelderblom et al., 1991 ). The ‘C’ series fumonisins lack the l-methyl group. Alternaria alternata lycopersici (AAL) toxins are structurally related compounds and share similar biological activity with fumonisin B, (Van Der Westhuizen et al., 1998; Wang et al., 1996). cHzcooH \C-CHZCHCOOH ('3 OH OH H i 3 Fumonisin B, oéc-CHZQHCOOH 011200011 on 0” 9“- Wow” FUMONlSln 32 on 0“ 0H . . Wow ' Fumomsm A1 on OH OH W Fumonisin C, CHa OR 01-13 0H NH2 on 0“ 9” W Alternaria toxin cnaoa OH, NH: ‘ [a - cocwcmcoonicnzcooI-q Figure 1.1. Structures of fumonisins and related mycotoxins. A. 5. Toxicokinetics To investigate the distribution and excretion of fumonisins, MC-labeled fumonisin B, was prepared by the addition of l4C-methionine to a culture of F usarium moniliforme and used in rats (N orred et al., 1993; Shephard et al., 1994a), swine (Prelusky et al. , 1994), and laying hens (Vudathala et al., 1994). In rats, fumonisins appear to be absorbed poorly and excreted rapidly (Norred et al., 1993; Shephard et al., 1994a; Shephard et al., 1995). In laying hens, fumonisin B, is poorly absorbed and is rapidly eliminated with the level of residues near negligible (Vudathala et al., 1994). In swine, fumonisin B, is primarily excreted in feces (~58%) and to a lesser extent in urine (~20%) (Prelusky et al., 1994). F umonisin B, is rapidly cleared from plasma and showed wide distribution with residues in liver > kidney > large intestine > brain > lung, heart and adrenal gland (Shephard et al. , 1992; Prelusky et al., 1994). In bile-cannulated pigs, bile recovery was ~71% with absence of Y phase, indicating enterohepatic circulation (Prelusky et al. , 1994). Results from both rat and pig studies indicate the possible enterohepatic circulation of fumonisin B,. Thus, it is possible that the true bioavailability of fumonisin B, was underestimated because bile cannulation may have interrupted normal enteroheptatic circulation of fumonisin B, and increased fecal excretion of the mycotoxin. Importantly, fumonisin residues accumulated in liver and to a lesser extent in kidney when pigs were fed fumonisins at dietary concentrations of 2-3 ppm for 24 days (Prelusky et al., 1996). Therefore, consumption of fumonisin-contaminated diets over extended periods may result in the accumulation of toxic residues in liver and kidney. 10 t A. 6. Metabolism Studies with MC-labeled fumonisin B, also showed that fumonisin B, is not metabolized by cytochrome P-450 monooxygenase or esterases, and is not a substrate for hepatic lipases (Cawood et al., 1994). However, partially hydrolyzed (hydrolysis of one of the ester groups) fumonisin B, was identified by mass spectrometry and nuclear magnetic resonance spectroscopy (Shepherd et al. , 1994c). A study focused on the structure-activity relationships of the fumonisins indicated that the intact molecule is responsible for the toxic and carcinogenic activity rather than a metabolite (Cawood et al., 1994; Gelderblom et al., 1993). A.7. Inhibition of Ceramide Synthase The chemical structure of fumonisin B, is strikingly similar to that of the sphingolipid sphinganine including the toxin’s long carbon chain (“backbone”), amine group at the C-2 position, and hydroxyl groups attached to the backbone (Figure1.2). Due to this relationship, fumonisin B, inhibits sphinganine (sphingosine) N-acyltransferase (ceramide synthase). Ceramide synthase is responsible for the acylation of sphinganine in the de novo biosynthetic pathway of sphingolipids as well as the re-acylation of sphingosine that is released from turnover of complex sphingolipids (Merrill et al., 1993a; Wang et al., 1991). The ICso for inhibition of ceramide synthase is ~0.1 1.1M and fumonisin B, is competitive with both sphinganine and the fatty acyl CoA, suggesting that portions of fumonisin B, occupy both the sphingoid base and the fatty acyl CoA binding sites on ceramide synthase (Wang et al., 1991). Inhibition of ceramide synthase by fumonisins causes accumulation of sphingoid ll OR Fumonisin B, 0H 0H CH, CH3 OR CH, OH NH2 R = cocmcmcoomCHzcoon Sphinganine OH WCHZOH NH2 Figure 1.2. Structures of fumonisin B, and sphinganine. 12 bases, especially, sphinganine. F umonisin-induced sphinganine accumulation was shown in the tissues of mice (Martinova and Merrill, 1995), rats (Kwon et al., 1997; Riley et al., 1994a), mink (Restum et al., 1995) and pigs (Riley et al., 1993a). Some of the sphinganine is metabolized to the l-phosphate and degraded to hexadecanal and ethanolamine phosphate, the latter being incorporated into phosphatidylethanolamine (Smith and Merrill, 1995). Sphinganine is also released from cells and appears in the blood and urine of rats (Riley et al., 1994a), ponies (Wang et al., 1992), mink (Morgan et al., 1997), pigs (Riley et al., 1993a) and monkeys (Shephard et al., 1996). The sphinganine/sphingosine ratio in blood (Wang et al., 1992; Riley et al., 1993a) and in urine (Morgan et al., 1997; Riley et al.,1994) has been proposed as a sensitive indicator of fumonisin exposure. Inhibition of ceramide synthesis by fumonisins also eventually depletes complex sphingolipids such as ceramide and sphingomyelin (Riley et al., 1993a; Wang et al., 1992). However, complex sphingolipids differentially respond to fumonisin B,. For example. sphingomyelin was more rapidly depleted than glycosphingolipids (Meivar Levy and Futerman, 1999; Merrill et al., 1993a). The fumonisin B,-induced alterations in complex sphingolipids may be responsible for various morphological changes seen in cell culture. Fumonisin B, disrupted axonal growth in cultured hippocampal neurons (Harel and Futerman, 1993) by affecting the formation or stabilization of axonal branches (Schwarz et al., 1995). Long-term culture with fumonisin B, affected fibroblast morphology and proliferation with a concomitant decrease in the synthesis of ganglioside GM3, the major glycosphingolipid in 3T3 fibroblasts and of sphingomyelin (Meivar Levy et al., 1997). All of the effects of fumonisin B, on cell morphology were reversed by addition of ganglioside GM3 even in the presence of fumonisin B, ; whereas, the bioactive intermediates sphinganine, l3 sphingosine, and ceramide had no effect. These data suggest that disruption of complex sphingolipid synthesis by fumonisin B, can impair the structural integrity of cells. B. SPHINGOLIPIDS B. 1. Introduction Sphingolipids are characterized by a common structural feature, i.e., a sphingoid base backbone such as D-erythro-l,3-dihydroxy-2-aminooctadec-4-ene (sphingosine). The “sphingosine” backbone was named by J. L. W. Thudichum in 1884 for its enigmatic (“Sphinx-like”) properties while he was studying the chemical constituents of brain (Thudichum, 1 884). Sphingolipids are found in all eucaryotic cells, where they are especially rich in the plasma membrane and the membranes of organelles responsible for their synthesis (endoplasmic reticulum and Golgi) and degradation (endosomes and lysosomes). The sphingolipids of mammalian tissues, lipoproteins, and milk include ceramides, sphingomyelins, cerebrosides, gangliosides and sulfatides. Sphingolipids are constituents of most foods. The sphingolipid content of pork, beef, and chicken is ~0.3 - 0.5 pmol/g (Blank et al., 1992). Whole milk, butter, and cheese contain ~0.5 - 1.0 pmol/g (Zeisel et al., 1986; Ahn and Schroeder, 2002). The sphingolipid content is relatively low in plant foods with the exception of soybean which contains ~2 pmol/g (Ohnishi and Fujino, 1982; Ahn and Schroeder, 2002). Dietary sphingolipids are hydrolyzed throughout the intestine (N ilsson, 1969; Schmelz et al. , 1994). Sphingolipids are involved in the signal transduction pathways that mediate cell growth, differentiation and cell death. Hydrolysis of sphingomyelin by sphingomyelinase generates ceramide as a cellular second messenger for tumor necrosis factor-0t (TNF-(x), IL- 14 l B, and other cytokines. Subsequently, ceramide can be hydrolyzed by multiple ceramidases to generate sphingosine in response to IL-IB (N ikolova-Karakashian er al., 1997) and may be further phosphorylated to form sphingosine l-phosphate in response to PDGF (Olivera et al., 1999a). Sphingosine and sphingosine l-phosphate modulate a variety of targets, mobilize cellular calcium, and regulate cell proliferation (Merrill eta]. , 1997a). Perturbation of sphingolipid metabolism by key enzyme inhibitors such as fumonisins appears to cause abnormal cell functions and certain diseases. B. 2. De nova Biosynthesis Sphingolipids are synthesized de novo from the amino acid serine and palmitoyl CoA (Figurel .3). In the endoplasmic reticulum, serine palmitoyltransferase condenses the two molecules to form 3-ketosphinganine which NADPH-dependent reductase rapidly reduces to form sphinganine, a free sphingoid base. Sphinganine is N—acylated with a long-chain fatty acid to form N-acyl-sphinganine (dihydroceramide) by ceramide synthase. Subsequently, a 4,5-trans double bond is introduced to form ceramide (N—acyl-sphingosine) (Merrill and Wang, 1992; Rother et al. , 1992). Additional reactions take place in the Golgi apparatus where head groups are added. For example, the transfer of a phosphorylcholine head group from phosphatidylcholine to the 1 position of ceramide yields sphingomyelin (Futerrnan and Pagano, 1991). Currently there are more than 300 known sphingolipids with distinct head groups (Bell et al. , 1993). All contain a long-chain (sphingoid) base backbone with D-erythra-sphingosine being the most prevalent backbone. However, there are more than 60 different sphingoid base backbones that vary in lengths of alkyl chain from 14 to 22 carbon atoms, degree of saturation and position of the double bonds, presence of a hydroxyl 15 0 COOH 20H Palmitoyl CoA Serine X0] + W H NH2 l SCoA i OH OH NH W l OH OPO3'CH2CH2N+(CH3) 3 NH WVWWVY sphingomyelin 0 Figure 1.3. De nova sphingolipid biosynthesis 16 group at position 4 and branching of the alkyl chain (Karlsson, 1970). Sphinganine and other intermediates in the de novo biosynthetic pathway of sphingolipids are highly bioactive and, under normal conditions, the cellular concentrations of these compounds are kept very low (Merrill et al., 1986). B. 3. Turnover The ordinary turnover of sphingolipids involves removal of the head groups such as phosphorylcholine from sphingomyelin by sphingomyelinases in the lysosomes (F igurel .4). Ceramidase catalyzes the cleavage of the amide-linked fatty acid of ceramide to form free sphingosine. Sphingosine can then be either reacylated to form ceramide or phosphorylated via sphingosine kinase to form sphingosine l-phosphate (Stoffel er al., 1968, 1970). Subsequently, sphingosine l-phosphate is cleaved to form ethanolamine phosphate and trans-2-hexadecanal by a pyridoxal phosphate-dependent lyase located in endoplasmic reticulum (Stoffel et al., 1968). The ethanolamine phosphate may be used for the synthesis of phosphatidylethanolamine and phosphatidylcholine. Another possibility for the metabolism of sphingosine has been suggested to involve methylation to form di- and trimethyl sphingosine (lgarashi et al., 1990; Hakomori and Igarashi, 1993). B. 4. Sphingolipid Signaling In addition to the normal turnover of sphingolipids. the metabolism of sphingolipids has evolved as an important metabolic pathway which plays essential roles in the production of cellular messengers to regulate cell behavior. That is, complex sphingolipids in membranes may also serve as substrates for agonist-induced signaling. The resulting 17 OH opo,-CH,CH2N+(CH3) 3 Sphingomyelin NH OH OH Ceramide WW OH OH Sphingosine kinasel NH2 OH OPO3H, Sphing°8ine 1W 1 NH, /\/\/\/\/\/\M + NH,CH,OPO,H, Hexadecanal O Ethanolamine phosphate Ether lipids Phosphatidylethanolamine Figure 1.4. Sphingolipid turnover 18 sphingolipid metabolites mediate a variety of cellular processes (Figure1.5). TNF-0t, interferon-Y, (Kim et al. , 1991 ) and 1a-25-dihydroxyvitamin D3 (Okazaki et al. , 1989, 1990) induces sphingomyelin hydrolysis by sphingomyelinase to produce ceramide which triggers monocytic differentiation of human premyelocytic leukemia HL-60 cells. A number of studies indicate that ceramide may regulate cell growth, either by inhibiting proliferation of normal fibroblasts (Hannun, 1994; Venable et al., 1994) or inducing proliferation of confluent, quiescent 3T3 fibroblasts (Olivera and Spiegel, 1993). Ceramide is a well-known mediator of programmed cell death or apoptosis, induced by TNF-oz, F as ligand, and ionizing radiation (Cifone et al., 1994; Haimovitz-Friedman et al. , 1994; Kolesnick and Golde, 1994; Obeid er al. , 1993; Hannun, 1994). TNF-0t is coupled to sphingomyelinase to produce ceramide (Kim et al., 1991), but the mechanism is not clearly understood. Exogenously added ceramide also induces apoptosis in various cell lines (Bielawska et al., 1997; Charles et al., 2001; Dawson et al., 1998; Hannun and Obeid, 1995; Jarvis et al., 1996; Saba et al., 1996). However, anti-apoptotic bcl-Z antagonizes ceramide action and prevents apoptosis (Pinton et al., 2001; Wieder et al., 1997; Zhang et al., 1996). Ceramide appears to act by affecting signaling molecules involved in the regulation of cell proliferation and apoptosis. Rafl is phosphorylated by a ceramide-activated protein kinase, activating the MAPK cascades (Raines et al., 1993; Yao et al., 1995). Also, a cytosolic ceramide-activated protein phosphatase (CAPP) has been identified as a molecular target for the action of ceramide (Dobrowsky and Hannun, 1992; Hannun, 1994). Okadaic acid, a phosphatase inhibitor, prevented ceramide-induced apoptosis and CAPP activity correlates with growth inhibition in numerous cell types (Hannun, 1994). Recently, ceramide was shown to induce dephosphorylation of Akt (protein kinase B) at serine 473, suggesting l9 Agonist\ Receptor Sphingomyeli Sphingomyelinase 1 OH OH _ R Growth inhibitory, Ceramide R cytotoxic/pro-apoptotic Ceramidase l O Sphingosine R OH OH Growth inhibitory. cytotoxic/pro-apoptotic Sphingosine l 2 kinase . . OH OPO H Sphmgosme 1-P R 3 2 Growth stimulatory. anti-apoptotic Figure 1.5. Sphingolipid signaling and cellular effects 20 activation of a phosphatase (Schubert et al. , 2000). Since Akt is involved in the activation of the cell proliferation pathway and regulation of apoptosis (Krasilnikov, 2000; Zhou et al. , 2000; Zundel and Giaccia, 1998), inhibition of Akt activity by ceramide-induced dephosphorylation may be one of the mechanisms by which ceramide inhibits growth and induces apoptosis. Some of the actions of ceramide in cellular processes appear to be mediated by deacylation to sphingosine which also is a putative second messenger. For example, IL-l B induces the hydrolysis of sphingomyelin to ceramide in primary cultures of rat hepatocytes and ceramide has been suggested to play a role in the down regulation of cytochrome p450 2C1 1 (CYP2C1 l ) and induction of a,-acid glycoprotein (AGP) (Chen et al. , 1995). Further, IL-l B increased not only sphingomyelinase activity, but also ceramidase with a concomitant increase of cellular sphingosine. Sphingosine was more potent than C2-ceramide in down- regulation of CYP2C1 l (N ikolova-Karakashian et aI. , 1997), suggesting that sphingosine is a mediator of the regulation of CYP2C 11 by IL-1 B. The sphingoid bases, sphingosine and sphinganine (Figure1.6), are produced from the metabolism of ceramide by ceramidase and during the de novo biosynthesis of sphingolipids, respectively (F igurel .3 and 1.4), and have received significant attention due to their inhibitory action on protein kinase C (PKC) (Hannun et al., 1986; Hannun and Bell, 1987) and PKC-dependent cellular processes in platelets (Hannun et al., 1986), neutrophils (Wilson et al., 1986), and HL-60 cells (Merrill et al., 1986). Sphingosine is competitive with diacylglycerol, phorbol dibutyrate, and Ca“, and also blocks PKC activation by unsaturated fatty acids and other lipids (el Touny et al., 1990; Oishi et al., 1988; Wilson et al., 1986). Like ceramide, sphingoid bases have been shown to be growth inhibitory (Merrill et 21 Sphinganine OH OH /\/\/\/\/\/\W NH, Sphingosine OH OH WW NH, Figure 1.6. Structures of sphingoid bases. 22 al., 1986; Chao et al., 1992; Dbaibo et al., 1995; Spiegel et al., 1993) and to induce apoptosis (Smyth et al., 1997; Dawson et al., 1998; Ohta et al., 1995; Shirahama et al., 1997; Jarvis et al., 1996; Sakakura et al., 1998). Sphingoid bases may cause cell death by affecting key signaling molecules that are involved in regulating cell growth and apoptosis. In addition to inhibiting PKC, sphingosine has been shown to inhibit phorbol dibutyrate binding (Merrill et al., 1986; Hannun et al., 1986), inhibit cell growth by arresting the cell cycle in the Go/G, phase (Chao et al., 1992; Dbaibo et al., 1995), activate the tumor suppressor, retinoblastoma protein, by inducing dephosphorylation (Pushkareva er al., 1995), inhibit p42/p44 MAPK activity (Sakakura et al. , 1998) and induce release of Ca++ from intracellular stores (probably via conversion to sphingosine 1-phosphate [SPP])(Ghosh et al., 1990). Recently, sphingosine also has been shown to dephosphorylate Akt during induction of hepatoma cell apoptosis (Chang et al. , 2001) (Figure 1.7). In addition, sphingosine suppresses expression of bcl-2 in human leukemic HL-60 cells (Sakakura et al., 1996) and bcl-x in human prostatic carcinoma DU-145 cells (Shirahama et al., 1997) during apoptosis. Addition of a phosphate group on sphingoid bases (sphingosine or sphinganine) by sphingosine kinase produces sphingoid base l-phosphates (SPP), the phosphorylated metabolites of sphingoid bases. In contrast to sphingoid bases, in most cases, SPP have been found to be growth stimulatory and anti-apoptotic (Conway et al., 1997; Olivera et al., 1999b; Spiegel, 1999). Platelet-derived growth factor (PDGF) stimulates sphingosine kinase activity and SPP acts as a second messenger in cell proliferation induced by PDGF and fetal calf serum (Olivera et al., 1999a; Olivera and Spiegel, 1993). Rapid activation of the Raf/ERK pathway (Wu et al., 1995) and increased DNA binding activity of activator protein- ] (AP-l) (Su et al., 1994) also may mediate the mitogenic effects of SPP, resulting in cell 23 Sphingoid bases /| Ras J\ lCaspase 9| . * Phosphate Group Figure 1.7. Effects of sphingolipids on cell proliferation signaling pathways 24 proliferation. SPP mediates nerve growth factor-induced neuronal survival and differentiation and has been shown to inhibit caspase activation during Fas- and ceramide- mediated apoptosis in Jurkat T lymphocytes (Cuvillier et al. , 1998). In rat periosteal RP-l 1 cells, SPP mediates the inhibition of apoptosis by forskolin-induced cAMP via stimulation of sphingosine kinase activity (Machwate et al., 1998). Recent work by Olivera et al. (1999b) demonstrated that overexpression of sphingosine kinase markedly increased the concentration of SPP in NIH 3T3 fibroblasts and HEK293 cells and is sufficient to promote growth in low-serum media, expedite the G,/S transition, and increase DNA synthesis. Furthermore, overexpression of sphingosine kinase in NIH 3T3 cells protected against ceramide-induced apoptosis (Olivera et al., 1999b). Since SPP can be produced from the metabolism of ceramide (Figurel.4), it has been suggested that the balance between intracellular ceramide and SPP may determine the cell fate (Cuvillier et al. , 1996). It should be noted, however, SPP inhibits proliferation of human hepatic ’myofibroblasts (Davaille et al., 2000). Therefore, SPP may be growth inhibitory and/or apoptotic under certain conditions. SPP may mediate the mitogenic effects of sphingoid bases. In Swiss 3T3 cells, sphingosine stimulated cellular proliferation via a protein kinase C-independent pathway (Zhang et al., 1990). Sphingosine induced a rapid increase in SPP levels in a concentration- dependent manner and there was a strong correlation with sphingosine’s effect on DNA synthesis (Zhang et al., 1991). When added together, there was no additive or synergistic effect, suggesting that sphingosine and SPP modulate cellular proliferation through a common pathway (Zhang et al., 1991). Both sphingosine and SPP potently mobilize calcium from internal stores with SPP acting more rapidly than sphingosine (Mattie et al., 1994; 25 Olivera et al., 1994; Zhang et al., 1991) and both activate phospholipase D (Desai et al., 1992), well known events in the control of cellular proliferation. B. 5. Disruption of Sphingolipid Metabolism as a Causative Factor in Fumonisin- induced Toxicity Fumonisins are cytotoxic and carcinogenic and this appears to be due to their cell- type specific effects on cellular behavior. Fumonisin B, is apoptotic in some cells while being mitogenic in others. The mechanism of action of fumonisins is not clearly understood but inhibition of ceramide synthase by fumonisins and the consequent depletion of complex sphingolipids and accumulation of sphinganine and metabolites may mediate the actions of fumonisins. Inhibition of acylation of sphinganine by fumonisins ultimately depletes complex sphingolipids and their role in fumonisin-induced toxicity is not clearly understood. Complex sphingolipids have been suggested to be involved in the regulation of cell growth and differentiation. Their profiles change as cells grow (Hakomori, 1981) and the majority of the glycosphingolipids and sphingomyelin are located in the external leaflet of the plasma membrane (Miller-Podraza et al., 1982), where they interact with cell surface receptors and mediate cell-cell, cell substratum, and cell-microbial interactions. This is further supported by results from in vitro studies which demonstrated that exogenously added gangliosides GM, or GM, inhibited cell growth by delaying the G, phase of the cell cycle (Laine and Hakomori, 1973) and can make cells refractory to growth stimulation by growth factors (Bremer and Hakomori, 1982). Also, addition of ganglioside GM3 induced HL-60 cells to differentiate into monocyte-like cells (Saito et al. , 1985). On the other hand, removal of ganglioside GM, by adding the B-subunit of cholera toxin or antibody which binds to this ganglioside is 26 mitogenic (Spiegel et al. , 1985) and potentiates the response to the growth factors (Spiegel and Fishman, 1987). Therefore, complex sphingolipids including gangliosides may play crucial roles in the regulation of cell growth and differentiation and their depletion during long-term exposure to fumonisins could contribute to carcinogenicity and other toxicities caused by fumonisins. Further investigations are needed to determine the role of depletion of complex sphingolipids in fiJmonisin-induced toxicities. The most well known consequence of fumonisin exposure is the accumulation of sphingoid bases which have been shown to inhibit cell growth and induce apoptosis (Figurel .5). Therefore, fumonisins may exert their toxicity via accumulation of sphinganine. The cytotoxic and apoptotic effects of sphingoid bases could be due to their inhibitory effect on protein kinase C (PKC), an important regulator of cell proliferation (Hannun and Bell, 1986). One of the major downstream targets of PKC is p44/p42 MAPK (or ERK) of which phosphorylation leads to phosphorylation and activation of activator protein-1 (AP-1). AP-l consists of dimers of fas and jun and regulates expression of genes involved in cell growth. Inhibition of ERK activity by sphingosine has been suggested to be a partial mechanism of action in sphingosine-induced apoptosis in various tumor cell lines (Sakakura et al., 1998). Therefore, fumonisin-induced sphinganine accumulation and down-regulation of ERK activity may contribute to fumonisin-induced cell death. Apoptosis often accompanies the suppression of anti-apoptotic signals such as PI3K/Akt (Jarvis et al. , 1997). The P13 K/Akt signaling pathway has been shown to mediate both Ras- and cytokine-induced protection from apoptosis. Activation of P13K recruits Akt to the plasma membrane where Akt is phosphorylated at serine 473 and threonine 308 (Stephens et al., 1998; Stokoe et al., 1997). Akt is a serine/threonine kinase involved in the 27 regulation of apoptosis via Bad and caspase-9 (Figure 1.8) (Krasilnikov, 2000; Zhou et al., 2000; Zundel and Giaccia, 1998). Phosphorylation of Akt activates its kinase function so that it may add a phosphate group on substrates such as the bcl-2 family protein Bad and caspase 9 (Krasilnikov, 2000). Phosphorylation of Bad serves as an anti-apoptosis signal by allowing bcl-x to inhibit release of cytochrome C from the mitochondria into the cytosol, which is a critical event in apoptosis. Also, phosphorylation of caspase 9 inhibits cleavage which is an activation process, preventing further activation of caspase 3, the executive protease protein in apoptosis. Recently, ceramide has been shown to dephosphorylate Akt and Bad during apoptosis (Basu et a1 ., 1998; Schubert et al., 2000). Inhibition of Akt activity by ceramide-induced dephosphorylation may be one of the mechanisms of ceramide action in its effect on growth inhibition and apoptosis induction. Importantly, sphingosine appears to mediate the action of ceramide on Akt because sphingosine also has been shown to reduce phosphorylation of Akt with concomitant reduction of its activity during sphingosine-induced apoptosis in human hepatoma cells (Chang et al., 2001). Moreover, expression of constitutively active Akt prevented sphingosine-induced apoptosis (Chang et al., 2001). Thus, fumonisin-induced sphinganine accumulation could induce Akt dephosphorylation and trigger cascades of molecular events to induce cell death. Sphingoid bases also may affect expression of genes involved in the regulation of cell survival and apoptosis. Numerous apoptotic agents depend on gene expression for their effects. Those genes in which expression has been documented to play a role in the apoptotic process include p53 tumor suppressor gene (Attardi et al., 1996; Bouvet et al., 1998), bcl-2 family member genes (Bargou et al., 1995a; Bargou et al., 1995b; Bruckheimer et al., 1998; Ealovega et al., 1996; Kondo et al., 1994), c-myc (Desbarats et al., 1996; Donzelli et al., 28 Sphingoid bases . Cytochrome C @® 0 Phosphate Group -B --11 ——» + -> Apoptosis Figure 1.8. Effects of sphingolipids on apoptotic signaling pathway 29 1999; Galderisi er al., 1999; Kihara Negishi et al., 1998; Wang et al., 1997), calmodulin (Dowd et al., 1991), heat-shock protein 70 (Garcia Bermejo et al., 1997), E 1 A (Putzer et al., 2000), and poly (ADP-ribose) polymerase (Putzer et al., 2000). Sphingoid bases suppress expression of bcl-2 (Sakakura et al., 1996) and bcl-x (Shirahama et al., 1997) which are involved in regulating apoptosis (Figure 1.8). Therefore, it is possible that fumonisin- induced sphinganine accumulation affects expression of these genes to induce apoptosis. Fumonisin B, was mitogenic and induced DNA synthesis in Swiss 3T3 cells (Schroeder et al., 1994). Fumonisin-induced mitogenicity was sphinganine-dependent because prevention of sphinganine accumulation using B-fluoroalanine inhibited fumonisin- induced DNA synthesis. Also, the exogenous addition of sphingoid bases stimulates DNA synthesis in Swiss 3T3 cells (Zhang et al., 1990), suggesting that fumonisin-induced DNA synthesis is due to sphinganine accumulation. Therefore, it appears that fumonisin-induced sphinganine accumulation has bimodal effects on cell growth depending on cell type. Both the apoptotic and proliferative processes may be regulated by a similar signaling pathway (Desbarats et al., 1996) and may involve a similar pattern of early gene responses (Grassilli et al., 1992). Therefore, fumonisin B, may affect expression of genes that are involved in both apoptosis and cell proliferation, inducing either cytotoxicity and carcinogenicity. C. LLC-PK, RENAL TUBULAR EPITHELIAL CELLS AS A MODEL LLC-PK, renal epithelial cells appear to be a good model system for the study of the mechanism of action of fumonisins. Kidney is the most sensitive organ to fumonisin- induced toxicity (Gumprecht et al., 1995; Riley et al., 1994a) with the primary target appearing to be the renal tubular epithelial cell. Fumonisin B, inhibited growth and 30 increased death in LLC-PK, cells with an early increase of intracellular sphinganine (Yoo et al., 1996). The dose of fumonisin B, which caused sphinganine accumulation was similar to the dose which caused toxicity, suggesting a possible relationship between sphinganine accumulation and fumonisin-induced toxicity (Yoo et al., 1992). D. HYPOTHESIS AND SPECIFIC AIMS This study is based on the hypothesis that fumonisin-induced kidney toxicity is mediated by disruption of sphingolipid metabolism. In this study, B-fluoroalanine, an inhibitor of serine palmitoyltransferase, will be used to dissect and differentiate the role of sphinganine accumulation and complex sphingolipid depletion in fumonisin-induced kidney toxicity. This study with LLC-PK, cells would help to better understand the role of disruption of sphingolipid metabolism in fumonisin-induced cytotoxicity and carcinogenicity in kidney and other tissues. The specific aims are: (1) Determine whether fumonisin B, kills cells by inducing apoptosis and evaluate the role of disruption of sphingolipid metabolism in fumonisin-induced cell death (2) Identify genes affected by fumonisin B, and investigate the molecular mechanism whereby fumonisin B, induces gene expression (3) Determine whether fumonisin B, affects signaling pathways involved in cell proliferation and cell death 31 CHAPTER II FUMONISIN B, INDUCES APOPTOSIS IN LLC-PK, RENAL EPITHELIAL CELLS VIA A SPHINGANINE- AND CALMODULIN-DEPENDENT PATHWAY 32 A. ABSTRACT Fumonisins are a family of mycotoxins produced by F usarium moniliforme, which is the most common mold found on corn throughout the world. These compounds are both toxic and carcinogenic for animals, and perhaps humans, with the kidney being the most sensitive organ to fumonisin toxicity. The molecular mechanism of fumonisin toxicity appears to involve disruption of de novo biosynthesis of sphingolipids and accumulation of sphinganine. The goals of this study were to determine whether fumonisin B, kills LLC-PK, renal kidney epithelial cells by inducing apoptosis and to identify genes affected by sphinganine that mediate fumonisin B,-induced cell death. F umonisin B, produced morphological changes (i.e., cell shrinkage, membrane blebbing) and time-dependent increases in DNA fragmentation demonstrating that the toxin induces apoptosis. Simultaneously, fumonisin B, blocked sphingolipid biosynthesis and caused accumulation of sphinganine. To further investigate the role of sphinganine in fumonisin B,-induced apoptosis, B-fluoroalanine (BFA) was used to inhibit serine palmitoyltransferase, which catalyzes an earlier step in the sphingolipid biosynthetic pathway. BFA blocked sphinganine accumulation and prevented fumonisin B,-induced DNA fragmentation, confirming that apoptosis induced by fumonisin B, is dependent upon accumulation of sphinganine. To examine gene expression, differential display reverse transcriptase polymerase chain reaction (DDRT-PCR) was applied to RNA isolated after 16 h of exposure to fumonisin B,. Differential expression in response to fumonisin B, of a gene identified as calmodulin has been verified by Northern analysis. Sphinganine appears to mediate the effect because BFA reduces induction of calmodulin mRNA by fumonisin B,. Fumonisin B, also increases calmodulin protein in a concentration-dependent manner and the calmodulin antagonist W7 33 blocks fumonisin B,-induced DNA fragmentation, supporting a role for calmodulin in fumonisin B,-induced apoptosis. In contrast, fumonisin B, has no effect on expression of bcl-2 family genes (box, bcl-Z, and bcl-x). These findings demonstrate that fumonisin B, kills LLC-PK, kidney cells by inducing apoptosis. Further, the results establish a sequence of events for fumonisin B,-induced apoptosis involving initial disruption of sphingolipid metabolism and accumulation of sphinganine (or a metabolite), which, in turn, induces expression of calmodulin. B. INTRODUCTION Fumonisins are a family of cytotoxic and carcinogenic mycotoxins produced by F usarium monilifome, one of the most common molds found on corn and other agricultural commodities in the United States and throughout the world (Marasas, 1996). Interest in the fumonisin mycotoxins arises, in part, because of their adverse effects on animal health. Fumonisins cause equine leukoencephalomalacia and hepatotoxicity (Marasas et al., 1988), porcine pulmonary edema (Harrison et al., 1990), and renal toxicity, hepatotoxicity, and liver cancer in rats (Gelderblom et al., 1988), with the kidney being the most sensitive organ to fumonisin toxicity (Bucci et al., 1998; Gumprecht et al., 1995). F umonisins also may pose a threat to human health, as consumption of contaminated maize has been correlated with esophageal cancer in areas of southern Afiica, China, and other countries (Marasas, 1996). Fumonisin toxicity appears to be mediated by disruption of sphingolipid metabolism (Merrill et al., 1 997 a). Fumonisins bear a remarkable structural resemblance to sphingosine and sphinganine, the long-chain (sphingoid) base backbones of more complex sphingolipids, and block de navo biosynthesis of sphingolipids by potently inhibiting sphingosine 34 (sphinganine) N—acyltransferase (Norred et al., 1992; Wang et al., 1991; Yoo et al., 1992). The inhibition causes depletion of more complex sphingolipids and accumulation of sphinganine (and sometimes sphingosine) in a variety of cultured cells (N orred et al., 1992; Wang et al.,1991; Yoo et al.,1992; Schroeder et al., 1994; Tolleson et al., 1999) and in the sera, liver, kidney, and urine of animals fed contaminated grains (Merrill et al., 1997b; Morgan et al., 1997; Riley et al., 1993a, 1994b; van der Westhuizen et al., 2001; Wang et al., 1992). Studies using a renal kidney epithelial cell line (LLC-PK,) have demonstrated a strong association between accumulation of sphinganine caused by fumonisin B, and inhibition of cell growth. morphological changes, and increased cell death (Y 00 et al., 1992, 1996). The mode by which fumonisin B, kills LLC-PK, kidney cells is not known; however, the mycotoxin induces apoptosis in other cell types and in viva (Lim et al., 1996; Tolleson et al., 1996a, 1996b, 1999; Sharma etal., 1997; Wang et al., 1996). Moreover, Tolleson et al. (1996a, 1996b, 1999) have shown that exogenous sphinganine induces apoptosis in human keratinocytes and that addition of B-haloalanine blocks fumonisin B,-induced sphinganine accumulation and DNA fragmentation, suggesting that accumulation of sphinganine plays a key role in apoptosis induced by fumonisin B,. Apoptosis induced by sphingosine in HL-60 cells is accompanied by a decrease in expression of bcl-2, an apoptosis suppresser gene (Sakakura et al., 1996). In addition, sphingosine induces apoptosis in androgen-independent human prostatic carcinoma DU-145 cells by down-regulating bcl-xL (Shirahama et al., 1997). Thus, sphingoid bases may mediate fumonisin-induced cell death by altering expression of a key gene(s). This study used LLC-PK, renal kidney epithelial cells as a model system to test the 35 hypothesis that fumonisin B, kills the cells by inducing apoptosis. Further, this study also tested the hypothesis that sphinganine accumulation caused by fumonisin B, mediates cell death by altering gene expression. Based on previous findings that sphingoid bases alter the expression of bcl-2 family genes, we have examined the effect of fumonisin B, on expression of members of this gene family. In addition, we have used differential display reverse transcriptase PCR as an unbiased approach to examine the influence of fumonisin B, on expression of other genes. C. MATERIALS AND METHODS C.l. Reagents and cell culture a-33P-dCTP was obtained from DuPont NEN and C20- sphinganine was obtained from Matreya Co. All other reagents were obtained from Sigma Chemical (St. Louis, MO) unless otherwise noted. Porcine kidney epithelial LLC-PK, cells were obtained from American Type Culture Collection (ATCC) and cultured in Dulbecco’s Modified Eagle Medium/F-12K nutrient solution (Gibco BRL) (1:1) with 5% fetal bovine serum (FBS), 100 units/ml penicillin, 100 jig/ml streptomycin sulfate, and 100 pM of L- serine (Gibco BRL). Cells were treated when they were 70-80 % confluent. C.2. Analysis of DNA fragmentation by agarose gel electrophoresis assay Frfirwig treatments, fragmented DNA from LLC-PKl cells was extracted as described (Sellins and Cohen, 1987). In brief, cells from 100-mm dish were harvested using rubber policeman and collected by centrifugation (5 min, 500 g). The medium was removed and the cell pellet was resuspended in phosphate-buffered saline (PBS). After another centrifuge, the pellet was suspended in 0.1 ml hypotonic lysis buffer (10 mM Tris, 10 mM EDTA, 0.5% Triton X-100, pH 8.0). Cells were incubated for 10 min at 4°C and the lysate was centrifuged (30 min, 36 13,000 g, 4 ° C). The supernatant containing fragmented DNA was treated with RNase A (0.4 pg/pl) for 1 h at 37°C and then incubated with proteinase K (0.4 lag/pl) for 1 h at 37°C. DNA was precipitated overnight by incubation at -20°C in 50% isopropanol and 0.5 M NaCl. The DNA precipitate was pelleted by centrifugation (30 min, 13,000 g, 4°C), dissolved in distilled H,O and then electrophoresed in a 2 % agarose gel containing ethidium bromide (0.5 jig/ml) in 0.5X Tris-borate-EDTA buffer. After electrophoresis, DNA was visualized with a UV transilluminator and photographs were taken using a Polaroid system. C.3. Flow cytometric analysis of apoptotic cells. Flow cytometric analysis was performed as described by Telford et al. (1994) with modifications. Briefly, treated cells were collected by trypsinization and resuspended into single cell suspension in culture media and fixed with 70% ethanol, reaching final concentration of 50-53%.. Cells were pelleted and resuspended in PBS (pH 7.4) containing 0.1% Triton X-100, 0.1 mM EDTA. After pelleting the cells again, cells were resuspended with DNA staining solution (50 pg/mL propidium iodide, 0.1% Triton X-100, 100 11M EDTA, and 0.05 mg/mL RNase A) and analyzed for the percentage of cells in the sub-GO/Gl using a F ACS Vantage Flow Cytometer (Becton Dickinson, San Jose, CA). C.4. Mass measurement of sphinganine The cellular concentration of sphinganine was determined using HPLC method (Merrill er al. , 1988). Briefly, cells were harvested in PBS and the unnatural sphingoid base, C20-sphinganine, was added as an internal standard. The sphingoid bases were extracted with chloroform and methanol and treated with base to remove interfering glycerolipids. After preparation of the o-phthalaldehyde derivatives, the long-chain bases were separated by reverse-phase HPLC using a C18 column and eluted isocratically with methanolz5mM potassium phosphate, pH 7.0 (90:10). 37 C.5. RNA isolation Total RNA was isolated by the method of Chomczynski and Sacchi (1987) using RNA STAT 60 (Tel-Test). The integrity and the equal amount of the RNA was verified by examining 28S and 18S bands on 1.0% formaldehyde-agarose gel containing ethidium bromide (0.5 pg/mL). C.6. Quantitation of bcl-2, bcl-x and bax mRNA levels Competitive RT-PCR was used to quantitate bcl-2, bcI-x and box mRNA levels (Wang et al, 1995). This method is based on a competitive PCR approach using non-homologous internal standards called PCR MIMICs with the use of the PCR-MIMIC construction kit (C lontech Laboratories, Palo Alto, CA). MIMICs are DNA fragments constructed for use in competitive PCR amplification for quantitation of target mRNA levels. Each PCR MIMIC consists of a heterologous DNA fragment with primer templates that are recognized by a pair of gene- specific primers. Thus, these templates “mimic” the target and are amplified during PCR. The procedure used one set of primers to amplify both target cDNA and an externally added MIMIC of known concentration. MIMICs for bcI-2. bcl-x and bar quantitation were constructed following the manufacturer’s direction. The primer pairs used for amplification of bcl-2. bcl-x and box mRNA. 1 pg of RNA was used routinely for cDNA synthesis (Clontech). After first-strand cDNA synthesis reaction, the cDNA was made into 100 pl final volume, and 4 pl were used for PCR. A typical PCR consisted of 0.2 pM dNTP, 2 pl MIMIC, 4 pl cDNA, 0.2 pM of each primer, PCR buffer, and Taq polymerase. The PCR profile was 95°C for 45 s, 60°C for 45 s, and 72°C for 2 min for 30 cycles, followed by 72 °C for 7 min. After PC R, aliquots of the reaction were analyzed on 1.8 % agarose gel (0.5 jig/ml ethidium bromide). The amount of mRNA was quantitated by comparing relative intensities of the amplified MIMIC and specific message bands. All results were normalized 38 against glyceraldehyde 3-phosphate dehydrogenase (G3PDH) mRNA quantitated with the use of identical PCR conditions. Results are expressed as level of expression relative to G3PDH. The primer pairs and competitor for G3PDH were purchased from Clontech. C.7. Densitometry Analysis Western and Northern blot images were subjected to densitometry analysis using Quantity One (Bio-Rad, Hercules, CA). C.8. Differential Display Reverse Transcription Polymerase Chain Reaction (DD RT- PCR) Differential display gel kits including RTase (Genhunter Corp., Nashville, TN) were used for this study. RNA from control and F umonisin B,-treated cells were isolated and used for RT reaction using poly T primer with A, G or C as outlined by the supplier. RT products were labeled with 25 pCi (ac-”S dATP and amplified with 15 different combinations using 3 different down stream primers and 5 different upstream primers in a thermocycler (GeneAmp, PCR System 2400, Perkin Elmer Norwalk, CT) in the presence of DNA polymerase (Perkin Elmer Norwalk, CT). Labeled cDNAs were fractionated by size on a 6% DNA sequencing gel at a constant 60 W at 50°C for 2 h. Gels were transferred to Whatman 3 filter paper and dried at 80°C for 1 h. The dried gel was exposed to X-ray film for 1-2 days to obtain an autoradiograph. The target cDNA was isolated from a dried gel and rearnplified in the absence of the radioactive dATP. These cDNAs were used for cloning and as probes for northern analyses. C.9. Northern analysis Northern analyses of total RNA was performed as previously described (Schroeder and Cousins, 1990) using PCR-amplified partial cDNAs labeled with 50 pCi 0t-32P-dCTP (DuPont NEN) by the random oligolabeling method (Prime-a-Gene Labeling System from Promega). C.l0. cDNA cloning After screening for the false positives, the PCR-amplified partial 39 cDNA was inserted into the pCR2.l-TOPO plasmid vector using TOPO TA Cloning Kit (Invitrogen). This plasmid was used to transform TOP 10 ONE shotTM Cells with One Shot kit (Invitrogen). C .1 1. DNA sequencing The transformed competent cells were propagated in LB-media and the plasmid was isolated (QIAprep Spin Miniprep Kit, QIAGEN, Germany) for sequencing at the Michigan State University Core Sequencing Facility. DNA sequences were obtained and submitted to the GenBank. C.12. Immunoblot analysis for calmodulin Cellular protein was extracted from LLC-PK, cells with lysis buffer (20 mM Tris-Cl, pH 8.0, 150 mM NaCl, 10 % glycerol, 1 % Triton X-100, 2 mM EDTA, 1 mM penylmethylsulfonyl fluoride, 10 BM of leupeptin and aprotinin cocktail solution), sonicated, and centrifuged. The supernatant protein was denatured by incubating for 5 min at 95 °C in Laemmli buffer. The sample was applied to 15 % SDS-bis acrylamide denaturing gel and transferred onto nitrocellulose membrane. The protein on the membrane was blocked by overnight incubation in 5 % non—fat dry milk/TBS- Tween 20 (0.1%), pH 7.4. The membrane was probed with anti—calmoduin antibody (Santa Cruz) raised against a goat calmodulin for 1 hour at room temperature and treated with anti- goat IgG conjugated with horseradish peroxidase for additional 1 hour at room temperature. Chemiluminescence solution from Santa Cruz was used to develop the immunoblot. C.13. Cell protein measurement. Total cell protein was determined by the method of Lowry et al. (1995) using bovine serum albumin as a standard. C.14. Statistical analysis. Data from sphinganine were subjected to ANOVA and Bonferroni’s all pairwise comparison tests. The Student t test was used to analyze flow cytometric data. Differences were considered statistically significant at p s 0.05 40 D. RESULTS LLC-PK, renal kidney epithelial cells were used as a model systemfor these studies because the kidney is the most sensitive organ to fumonisin toxicity (Bucci et al., 1998; Gumprecht et al., 1995) and Y00 et al. (1992, 1996) have shown a strong association between death of LLC-PK, cells caused by fumonisin B, and accumulation of sphinganine. F umonisin B, induces apoptosis and causes sphinganine accumulation in LLC-PK, cells. Morphological changes characteristic of apoptosis, including cell shrinkage, membrane blebbing, chromatin condensation, and the formation of apoptotic bodies were observed in LLC-PK, cells cultured with 50 [1M fumonisin B, (data not shown). To further investigate the mechanism of death, LLC-PK, cells cultured with fumonisin B, were examined for the presence of DNA fragmentation, a hallmark of apoptosis. Control cultures (without toxin) showed no evidence of DNA fragmentation for up to 48 h (data not shown). In contrast, fumonisin B, produced time-dependent increases in a 200—bp DNA ladder indicative of apoptosis (Figure 2.1A). DNA fragmentation was detectable within 24 h of exposure to fumonisin B,, with a maximal response at 48 h. Fumonisin B, increased the percentage of cells undergoing apoptosis by about ~6.5-fold at 48 b (Figure 2.1B). Fumonisin B, at 50 11M also produced time-dependent increases in sphinganine. Fumonisin B, caused sphinganine accumulation within 6 h (Figure 2.2). Maximal accumulation of ~1 .7 nmol/mg protein (~50-fold that of controls) was achieved by 48 h. Thereafter, sphinganine in fumonisin-treated cultures decreased to ~12 nmol/mg protein by 72 h. fl-Fluaroalanine blocks sphinganine accumulation caused by fumonisin B,. The findings demonstrate that fumonisin B , induces kidney cell apoptosis over a time course that is consistent with dependence upon accumulation of sphinganine but do not prove a cause 41 -..:S‘JI rp'm— Culture Period (h) M 0 3 616243648 Apoptotlc Cells (Fold Increase) O H N DJ A Ur O\ \l 00 Control F B, Figure 2.1. Fumonisin B, induces internucleosomal DNA fragmentation in LLC-PK, kidney cells. (A) LLC-PK, cells were cultured in the presence of fumonisin B, at 50 11M for various lengths of time and DNA fragmentation was assessed by agarose gel electrophoresis as described under Materials and Methods Results are from a representative experiment. M: 123—bp DNA markers. (B) LLC-PK, cells were cultured in the absence or presence of fumonisin B, at 50 11M and the percentage of cells in the sub-GO/Gl apoptotic region was determined by flow cytometric analysis as described under Materials and Methods. 42 2000 1500 - ...x U" o O O O O I +fl Sphlngamne (pmol/mg protein) C _500 1 1 1 1 1 1 1 1 0102 304 50607080 Culture Period (h) Figure 2.2. Fumonisin B, causes sphinganine accumulation in LLC-PK, kidney cells. LLC-PK, cells were cultured either in the absence (0) or presence (0) of fumonisin B, at 50 11M for various lengths of time. Cellular concentrations of sphinganine were determined by HPLC as described under Materials and Method. Each value is the mean :1: SD (n=3) from a representative experiment. Error bars that are not visible are hidden by the symbols. * value is significantly different (p 50.05) from corresponding control value. 43 and effect relationship. To obtain more information about the mechanism, studies were conducted using B-fluoroalanine. B-haloalanines act as mechanism-based irreversible inhibitors of serine palmitoyltransferase by substituting for the substrate serine during Schiff base formation (Medlock and Merrill, 1988). B-fluoroalanine is ~1000-fold more potent than B-chloroalanine for inhibition of serine palmitoyltransferase. Because serine palmitoyltransferase catalyzes the initial, rate-limiting step of sphingolipid biosynthesis, B- haloalanines block de navo synthesis of sphingolipids but would not be expected to cause accumulation of sphinganine. LLC-PK, cells were cultured with combinations of fumonisin B, (50 11M) and B- fluoroalanine (50 BM) to determine the effects on sphinganine accumulation and DNA fragmentation. The addition of fumonisin B, alone increased sphinganine (Figure 2.3) and stimulated DNA fragmentation (Figure 2.4); whereas, B-fluoroalanine alone had no effect on sphinganine (Figure 2.3) or DNA fragmentation (Figure 2.4). In contrast, the stimulatory effects of fumonisin B, on sphinganine accumulation and DNA fragmentation were completely blocked when B-fluoroalanine was added together with fumonisin B, (Figures 2.3 and 2.4). These results establish that the de nova biosynthesis (and, presumably, accumulation) of sphinganine or a metabolite such as the l-phosphate is required for the stimulation of DNA fragmentation by fumonisin B,. F umonisin B, does not affect expression of box. bc1-2, and bcl-x. Since altered expression of bcl-2 family genes has been implicated in apoptosis induced by free sphingoid bases, the effects of fumonisin B, on mRN A of these genes was examined using competitive RT-PCR. Figure 2.5 shows bax, bcl-Z, and bcI-x mRNA after 16 h of exposure to an apoptotic concentration of fumonisin B, (50 |J.M). Densitometric analysis indicated the toxin 44 1200 ... 1000 . , '5. 800 .. . e i rte 4"" “ a \9 200 e b 0 i b b -_J Control FB, BFA FB,+ BFA Treatment Figure 2.3. Inhibition of fumonisin B,-stimu1ated sphinganine accumulation in LLC- PK, cells by B-fluoroalanine. LLC-PK, cells were cultured in the absence or presence of fumonisin B, (FB,) at 50 11M and B—fluoroalanine (BFA) at 50 11M as indicated. After 16 hours, cellular concentrations of sphinganine were determined by HPLC as described under Materials and Methods. Each value is the mean :1: SD from a representative experiment. Means with differing letters are significantly different (p $0.05). 45 Figure 2.4. Inhibition of fumonisin B,-induced DNA fragmentation in LLC-PK, cells by B—fluoroalanine. LLC-PK, cells were cultured in the absence or presence of fumonisin B, (FB,) at 50 pM and/or B—fluoroalanine (BFA) at 50 uM. After 72 hours, cells were harvested and DNA fragmentation was assessed by agarose gel electrophoresis as described under “Materials and Methods”. Results are from a representative experiment. M: 123 base- pair DNA markers. 46 had no effect on expression of these genes. These results suggest that alterations in bcl-2 family gene expression are not necessary for fumonisin B,-induced apoptosis in LLC-PK, kidney cells. F umanisin B, induces expression of calmodulin mRNA by a sphinganine-dependent pathway. As an alternative approach to studying individual mRNAs, we have further examined RNA isolated after 16 h of culture with 50 BM fumonisin B, using DDRT-PCR. This method provides an unbiased approach to analyze changes in gene expression associated with a treatment without DNA sequencing information (Liang and Pardee, 1992). Using DDRT-PCR we have identified ~40 cDNAs that are differentially displayed in response to fumonisin B,. Figure 2.6 shows examples of three up-regulated kidney cell cDNAs (K1 from control cells and K2 and K3 from cells cultured with fumonisin B,. Northern analysis using the amplified cDNAs as probes confirmed differential expression of K3 (Figure 2.7); whereas, K1 and K2 were false positives (data not shown). The sequence of the K3 insert was determined and submitted to GenBank. The highest homology corresponds to human and mouse calmodulin, with 98% base sequence identity and 100% amino acid identity to human calmodulin (Figure 2.8). These findings establish that calmodulin mRNA is increased in LLC-PK, cells after 16 h of exposure to fumonisin B, and are consistent with, but not proven, to be dependent upon prior accumulation of sphinganine. To obtain information about the role of sphinganine in fumonisin-induced calmodulin expression, B-fluoroalanine was used to block sphinganine accumulation. As shown in Figures 2.9A and 2.9B, the addition of fumonisin B, alone stimulated calmodulin mRNA ~6.5-fold. In contrast, the stimulatory effect of fumonisin B, on calmodulin expression was reduced by about 50% when B-fluoroalanine 47 Figure 2.5. Fumonisin B, does not affect gene expressions of bcl-2 family genes in LLC- PK, cells. LLC-PK, cells were cultured in the absence (C) or presence (F) of fumonisin B, at 50 BM. After 16 hr, total RNA were isolated and box. bcl-Z. and bcI-x mRNA was quantitated by RT-PCR as described under Materials and Methods. G3 PDH, glyceraldehyde 3-phosphate dehydrogenase. Results are from a representative experiment. 48 K3 K1 K2 Figure 2.6. Differential display of LLC-PK, kidney cell cDNAs. LLC-PK, cells were cultured in the absence (C) or in the presence (F) of fumonisin B, at 50 11M. After 16 hours, total RNA was isolated and used for DDRT-PCR as described under Materials and Methods. The two primer combinations that were used for this experiment were: 5'-Tl 1-G-3' and 5'- GCAATCGATG-3’ (lanes 1 and 2) and 5'-T1 1-A-3' and 5'-CCGAAGGAAT-3‘ (lanes 3 and 4). 49 Figure 2.7. Fumonisin B, induces expression of K3 mRN A. LLC-PK, cells were cultured in the absence (C) or presence of fumonisin B, (F B,) at 50 uM. After 16 hr, total RNA was isolated and examined via Northern analysis as described under Materials and Methods. 50 K3 tgtgtttgataaggatggcaatggctatattagtgcagcagagcttcgccatgtgatgac 60 Human CaM tgtgtttgataaggatggcaatggctatattagtgctgcagaacttcgccatgtgatgac 400 K3 aaaccttggagagaagttaacagatgaagaggttgatgaaatgatcagggaagcagatat 120 Human CaM aaaccttggagagaagttaacagatgaagaagttgatgaaatgatcagggaagcagatat 460 K3 tgatggtgatggtcaagtaaactatgaagagtttgtacaaatgatgacagcaaagtgaag 180 Human CaM tgatggtgatggtcaagtaaactatgaagagtttgtacaaatgatgacagcaaagtgaag 520 K3 acgttgtacagaatgtgttaaatttcttgtacaaaattgtttatttgccttttctttgtt 240 Human CaM accttgtacagaatgtgttaaatttcttgtacaaaattgtttatttgccttttctttgtt 580 Figure 2.8. Sequence homology of K3 DNA insert with human calmodulin. The K3 cDNA insert was cloned and sequenced as described under Materials and Methods and submitted to GenBank for identification. Shown are the K3 cDNA insert and the corresponding sequence of human calmodulin (human CaM). K3 cDNA bases shown in bold are different from the corresponding base of human calmodulin. 51 F;-‘_‘ i I-- - J. CaM mRNA (Fold Stimulation) O r-‘ N DJ A U! ON \1 Con F B, BFA+FB, Figure 2.9. Inhibition of fumonisin B,-induced calmodulin expression in LLC-PK, cells by B-fluoroalanine. LLC-PK, cells were cultured in the absence or presence of fumonisin B, (FB,) at 50 pM and/or B—fluoroalanine (BFA) at 50 11M. After 16 hr, total RNA was isolated. (A) Northern analysis using K3 cDNA as a probe as described under Materials and Methods. (B) Densitometric analysis of the Northern blot as described under Materials and Methods. Results are from a representative experiment. 52 was added together with fumonisin B,. This finding establishes that fumonisin B,-induced expression ofcalmodulin mRNA is dependent on accumulation of sphinganine. F umonisin B, increases calmodulin protein in LL C -PK , kidney cells. Additional studies were conducted to examine the effect of fumonisin B, on calmodulin protein. LLC- PK, cells were cultured with fumonisin B, for 16 h and calmodulin protein was assessed by Western analysis. Fumonisin B, produced concentration-dependent increases in calmodulin protein with a maximun response of ~4-fold at 50 BM (the highest dose tested) (Figures 2.10A and 2.10B) Calmodulin antagonist bIocksfumonisin-induced DNA fragmentation. T h e findings to this point establish that fumonisin B, induces calmodulin expression as well as apoptosis in LLC-PK, cells via a sphinganine-dependent mechanism. To determine whether calmodulin plays a causal role in fumonisin-induced apoptosis, studies were conducted using W7. W7 acts as a calmodulin antagonist by competitively binding to calmodulin at the same site to which calmodulin-dependent enzymes bind (Osawa et al., 1998). LLC-PK, cells were cultured with combinations of fumonisin B, and W7 to determine their effects on DNA fragmentation. The addition of fumonisin B, alone increased DNA fragmentation, while W7 alone at either 10 or 20 BM had no effect (Figure 2.1 1). However, the stimulatory effect of fumonisin B, on DNA fragmentation was almost completely eliminated when W7 was added together with fumonisin B,. These results establish that fumonisin B, induces apoptosis in LLC-PK, kidney cells by a calmodulin-dependent mechanism. E. DISCUSSION Fumonisin mycotoxins are of growing concern because of their prevalence in com 53 FB, 0 10 25 B c 5 r ~—_——_—~ ~_#_ ..__-____ L l '32 8 4 _ O '5 a: .2 3 3 __ E .2 e a 2» "D g a I - U v Fumonisin B, (uM) Figure 2.10. Concentration of fumonisin B, induction of calmodulin protein in LLC-PK, kidney cells. LLC-PK, cells were cultured with various concentrations of fumonisin B, (F B,). After 16 hr, cellular protein was isolated. (A) Immunoblot analysis of calmodulin as described under Materials and Methods. (B) Densitometric analysis of the immunoblot as described under Materials and Methods. Results are from a representative experiment. 54 Figure 2.11. Inhibition of fumonisin-induced DNA fragmentation in LLC-PK, cells by calmodulin antagonist, W7. LLC-PK, cells were cultured in the absence or presence of fumonisin B, as indicated. W7 was also added to some cultures 16 hr after the addition of fumonisin B,. After 48 hr, DNA fragmentation was assessed by agarose gel electrophoresis as described under Materials and Methods. Results are from a representative experiment. 55 and corn-based products (Shephard et al.. 1996) as well as other agricultural commodities and because of their toxic and carcinogenic properties for animals and, perhaps, humans (Marasas, 1996). Across species, the most sensitive organ to fumonisin toxicity is the kidney (Bucci et al, 1998; Gumprecht et al, 1995). The primary response of kidney to consumption of feed containing fumonisin B, is tubular epithelial cell apoptosis (Bucci et al., 1998; Gumprecht et al, 1995). Yoo et al. (1992, 1996) have reported that fumonisin B, is cytotoxic for LLC-PK, porcine renal tubular epithelial cells, suggesting that this cell line may be an appropriate model to further investigate the molecular mechanism of renal toxicity caused by fumonisins. A major finding of the current study is that fumonisin B, kills LLC-PK, cells by inducing apoptosis. The concentration of fumonisin necessary to induce apoptosis, 50 pM, is identical to that previously reported to be toxic for these cells (Yoo et al., 1992, 1996) and is similar to that which induces apoptosis in a variety of other cultured cells, including human keratinocytes, fibroblasts, esophageal epithelial cells, and hepatoma cells (Tolleson et al., 1996b). Numerous studies have demonstrated that the mechanism of action of fumonisins appears to involve disruption of sphingolipid metabolism via inhibition of sphinganine N- acyltransferase (Merrill et al., 1997a). The inhibition causes accumulation of sphingoid bases in a variety of cell types (Norred et al., 1992; Wang et al., 1991; Yoo et al., 1992; Schroeder et al., 1994; Tolleson et al., 1999) and in the tissues and urine of animals fed fumonisin-contaminated grain (Merrill et al., 1997b; Morgan et al., 1997; Riley et al., 1993a, 1994; van der Westhuizen et al., 2001; Wang et al., 1992). In the present study, an apoptotic concentration of fumonisin B, caused cellular sphinganine to increase to ~1.0 nmol/mg of protein at least 8 h prior to evidence of DNA fragmentation. There are several possible 56 p. - .lfl‘hmrl‘ mechanisms whereby disruption of sphingolipid metabolism could cause apoptosis. First, fumonisin B, might block the formation of a complex sphingolipid(s) that is required for growth and cell survival. Hanada et al. (1992) showed in mutant Chinese hamster ovary cells (which can not synthesize complex sphingolipids) that addition of exogenous complex sphingolipids restored the ability of the cells to proliferate. Moreover, the addition of ceramide derivative reversed fumonisin inhibition of axonal growth in hippocampal neurons (Harel and Futtemlan, 1993) and the addition of N-acetylsphinganine partially protected human keratinocytes from fumonisin B,-induced apoptosis (Tolleson et al., 1999). Moreover, Yoo et al. (1996) demonstrated that, when B-chloroalanine was used to inhibit de nova synthesis of sphingolipids in LLC-PK, cells to a similar degree as fumonisin B,, it both inhibited cell growth and increased cell death. In contrast, when de nova synthesis of sphingolipids was blocked in the present study using a more specific inhibitor of serine palmitoyltransferase, B-fluoroalanine, there was no effect on DNA fragmentation. Thus, reduced synthesis of complex sphingolipids does not appear to explain the apoptotic effect of fumonisin B, in LLC-PK, cells. The second possible mechanism whereby disruption of sphingolipid metabolism could mediate fumonisin B, induction of apoptosis is by causing accumulation of the sphingoid base sphinganine (or a metabolite). Free sphingoid bases have been shown to be both growth inhibitory and cytotoxic for a variety of cell systems (Hannun et al., 1991; Stevens et al., 1990). Moreover, Yoo et al. (1996) reported that sphingoid base accumulation was, in part, responsible for fumonisin toxicity in LLC-PK, kidney cells. In support of a causal relationship between sphinganine accumulation and apoptosis, our study shows that fumonisin B,-induced DNA fragmentation was virtually eliminated upon the 57 addition of B-fluoroalanine, which blocked sphinganine synthesis and reduced the cellular sphinganine to a concentration similar to that of control cultures. These results provide compelling evidence that fumonisin B, induces apoptosis in LLC-PK, kidney cells by a pathway that involves disruption of sphingolipid metabolism and requires accumulation of sphinganine. Sphinganine could mediate the apoptosis induced by fumonisin B, either by a direct action or via conversion to a metabolite. Exogenous sphingoid bases can be converted to ceramide, a sphingolipid that induces apoptosis (Obeid et al., 1993). However, a role for ceramide in mediating fumonisin-induced apoptosis seems unlikely because the conversion of sphinganine to ceramide is blocked by fumonisins. Another sphinganine metabolite that could mediate apoptosis is the l-phosphate metabolite. Though sphingoid base l-phosphate is most often considered to be mitogenic (Zhang et al., 1991), recent studies have demonstrated that this molecule also has antiproliferative properties (Bomfeldt et al., 1995; Davaille et al., 2000). The mechanism by which fumonisins induce apoptosis via sphinganine is not clear; however, recent studies demonstrate that apoptosis induced by sphingosine in HL-60 cells is accompanied by a decrease in expression of bcl-2, an apoptosis suppresser gene (Sakakura et al., 1996). In addition, sphingosine induces apoptosis in androgen-independent human prostatic carcinoma DU-45 cells by down-regulating bcl-x, (Shirahama et al., 1997). Therefore, sphingoid bases may mediate fumonisin-induced cell death by regulating expression of a bcl-2 family gene(s). However, in our study, an apoptotic concentration of fumonisin B, had no effect on box, bcl-2, or bcI-x mRNA after 16 h of exposure. These results indicate that apoptosis of LLC-PK, cells is not due to an effect on bcl-Z family gene 58 expression. As an alternative to studying individual mRNAs, the present study used DDRT- PCR as an unbiased approach to analyze changes in gene expression associated with fumonisin B, treatment. A major finding of our study using this technique is that fumonisin B, induces calmodulin expression. The mycotoxin induces calmodulin expression in a concentration-dependent manner with maximal effects of ~6.5-fold at the mRN A level and ~4—fold at the protein level with an apoptotic concentration of fumonisin B,. Fumonisin B, could be envisioned to increase calmodulin expression either via a mechanism that is independent of its effects on sphingolipid metabolism or by a mechanism that is mediated via disruption of sphingolipid metabolism. In support of a causal role of sphinganine, fumonisin B, stimulation of calmodulin expression was reduced by ~50% upon the addition of B-fluoroalanine, which blocked sphinganine synthesis. This result provides convincing evidence that fumonisin B, induces calmodulin expression by a pathway that requires accumulation of sphinganine. The mechanism by which fumonisins stimulate calmodulin expression via sphinganine is not known. Transcriptional regulation of calmodulin is still poorly understood. However, Toutenhoofd et al. (1998) have identified possible AP-l binding sites and TATA-like sequences 27 nucleotides upstream of the transcriptional start site in the 5' flanking sequence of human calmodulin 2 and the sphingoid base sphingosine has been shown to induce DNA binding activity of AP-l (Su et al., 1994). In addition, Wang et al. (1996) have hypothesized that fumonisin-induced growth arrest and apoptosis in African green monkey kidney cells was a result of AP-l repression via inhibition of protein kinase C. No attempt was made to relate changes to fumonisin disruption of sphingolipid metabolism; however, the effects on protein kinase C are consistent with fumonisin-induced 59 sphinganine accumulation, since sphingoid bases are well-established inhibitors of protein kinase C (Hannun et al., 1991). In support of a causal role of calmodulin induction in apoptosis, another major finding of our study is that DNA fragmentation induced by fumonisin B, was virtually eliminated upon the addition of the calmodulin antagonist W-7. This finding provides strong evidence that fumonisin B, induces apoptosis in LLC-PK, kidney cells by a pathway that requires calmodulin. This discovery is consistent with the results of several recent studies that indicate a role for calmodulin in apoptosis. For example, Rosenthal et al. (1998) have reported that W-7 blocked apoptosis induced by sulfur mustard. In addition, Micoli et al. (2000) have demonstrated that the proapoptotic effect of HIV-1 envelope glycoprotein gp160 is blocked by the calmodulin antagonist trifluoroperazine. The mechanism by which fumonisin induction of calmodulin leads to apoptosis has not been examined but could involve activation of calcineurin (Stemmer and Klee, 1994), which, in turn, has been shown to dephosphorylate BAD, a proapoptotic member of the bcl-2 family (Wang et al.,'1999). Activation of calcineurin by calmodulin is achieved by lowering the threshold concentration of calcium needed for activation and calcium mobilizing agents such as ionomycin and A23 1 87 induce dephosphorylation of BAD, indicating activation of calcineurin (Wang et al., 1999). Since fumonisin B, has been shown to increase sphinganine l-phosphate in some cell types (Smith and Merrill, 1995) and sphingosine 1-phosphate mobilizes cellular calcium (Olivera et al., 1994), fumonisins may trigger bimodal regulation of calcineurin to dephosphorylate BAD and induce apoptosis. Taken together, the findings of this study using LLC-PK, cells provide a plausible molecular mechanism to explain toxicity of fumonisins to kidney renal tubular epithelial 60 cells. Fumonisin B, was found to disrupt sphingolipid metabolism and cause apoptosis by a pathway that involved initial accumulation of sphinganine, which, in turn, stimulated induction of calmodulin. 61 CHAPTER III SPHINGANINE l-PHOSPHATE MEDIATES CALMODULIN EXPRESSION INDUCED BY FUMONISIN B, IN LLC-PK, PORCINE RENAL TUBULAR EPITHELIAL CELLS 62 A. ABSTRACT Fumonisins are a family of mycotoxins produced by F usarium moniliforme, which is the most common mold found on corn throughout the world. These compounds are both toxic and carcinogenic for animals, and perhaps humans, with the kidney being the most sensitive organ to fumonisin toxicity. Fumonisin B, , the most prevalent of these mycotoxins, induces calmodulin expression in LLC-PK, porcine kidney epithelial cells and causes apoptosis via a calmodulin-dependent pathway (Kim et al., 2001). The molecular mechanism of calmodulin induction by fumonisin B, is not known, but appears to involve disruption of sphingolipid metabolism. Fumonisin B, inhibits ceramide synthase in the de nova biosynthetic pathway of sphingolipids and causes depletion of complex sphingolipids and accumulation of sphingoid bases. In the present study, LLC-PK, porcine renal tubular epithelial cells were used to assess the influence of fumonisin B, on calmodulin mRNA stability and to determine whether induction of calmodulin by fumonisin B, is mediated by the depletion of complex sphingolipids or accumulation of the sphingoid base sphinganine or the l-phosphate metabolite. Fumonisin B, increased both calmodulin mRNA and calmodulin protein in concentration-dependent manners, but did not influence stability of calmodulin mRNA, suggesting that fumonisin B, induces calmodulin by increasing transcription. Moreover, the transcription inhibitor actinomycin D blocked fumonisin- induced apoptosis. Neither the addition of exogenous sphinganine nor B-fluoroalanine (an inhibitor of serine palmitoyltransferase which depletes complex sphingolipids without causing sphinganine accumulation) induced calmodulin expression; whereas, sphinganine 1-phosphate increased calmodulin mRNA. Taken together, these findings suggest that fumonisin B, induces calmodulin in LLC-PK, kidney cells by disrupting sphingolipid 63 metabolism and causing accumulation of the sphinganine metabolite sphinganine 1- phosphate which, in turn, increases transcription of the calmodulin gene. B. INTRODUCTION Fumonisins are toxic and carcinogenic toxins produced by F usarium moniliforme, the most common mold found on corn in the United States and throughout the world (Marasas, 1996). F umonisin B, is the most abundant of these mycotoxins and has been correlated with several animal diseases with kidney being the most sensitive organ (Bondy et al., 1996; Bucci et al., 1998). In rats, consumption of feed contaminated with fumonisin B, causes apoptosis in renal tubular epithelial cells within 48 hr (Voss et al, 2001) and renal tubular adenomas and carcinomas within two years (Howard et al., 2001). In addition to kidney carcinogenicity, fumonisin B, consumption has been linked to esophageal (Marasas et al. , 1988), liver (Ueno et al. , 1997) and stomach (Groves et al. , 1999) cancers in humans. The molecular mechanism of fumonisin-induced apoptosis is not clear, but appears to involve calmodulin induction (Kim et al. , 2001). In LLC-PK, renal tubular epithelial cells, fumonisin B, induces calmodulin expression and calmodulin activity is required for fumonisin B,-induced apoptosis (Kim et al., 2001). The mechanism by which fumonisin B, induces calmodulin is not clear. Due to structural similarity with sphingoid bases, fumonisin B, inhibits ceramide synthase (sphinganine N-acyltransferase) (Figure 3.1). The inhibition causes depletion of complex sphingolipids as well as accumulation of sphingoid bases and/or the l-phosphate metabolite (Smith and Merrill, 1995; Norred et al., 1992; Wang et al., 1991; Yoo et al., 1992; Schroeder et al., 1994; Tolleson et al., 1999) and in the tissues and urine of animals fed fumonisin- 64 Serine + Palmitoyl CoA B—fluoroalanine ——-llSerine palmitoyltransferase 3-ketosphinganine Sphinganine —> Sphinganine l-phosphate Fumonisin Bl ll Ceramide synthase Ceramide Complex Sphingolipids Figure 3.1. Be nova biosynthetic pathway of sphingolipids and inhibition of ceramide synthase by fumonisin B, 65 contaminated grain (Merrill et al.. 1997b; Morgan et al., 1997; Riley et al., 1993a, 1994; van der Westhuizen et al., 2001; Wang et al., 1992). Fumonisin-induced toxicities both in vitro and in viva are dependent on disruption of sphingolipid metabolism (Martinova and Merrill, 1995; Restum et al., 1995; Yoo et al., 1996; Yoo et al., 1992) and fumonisin-induced calmodulin expression in LLC-PK, porcine renal epithelial cells was blunted by B- fluoroalanine, an inhibitor of de novo biosynthesis of sphingolipids which blocks at an earlier step than fumonisins and prevents sphinganine accumulation (Kim et al., 2001). Therefore, fumonisin-induced calmodulin expression is, at least in part, dependent on accumulation of sphinganine and/or a metabolite such as the l-phosphate (Kim et al., 2001). In the present study, LLC-PK, porcine renal tubular epithelial cells were used to test the hypothesis that the induction of calmodulin mRNA by fumonisin B, is mediated by the depletion of complex sphingolipids or accumulation of sphinganine and/or sphinganine 1- phosphate. In addition, the transcription inhibitor actinomycin D was used to examine the effect of fumonisin B, on the stability of calmodulin mRNA as well as the requirement for novel gene transcription in fumonisin-induced apoptosis. C. MATERIALS AND METHODS C.1. LLC-PK, cell culture Porcine kidney epithelial LLC-PK, cells were obtained from American Type Culture Collection (ATCC) and cultured in Dulbecco’s Modified Eagle Medium/F -12K nutrient solution (Gibco BRL) (1:1) containing 5% (v/v) fetal bovine serum (FBS), 100 units/ml penicillin, 100 [lg/ml streptomycin sulfate, and 100 11M of L-serine (Gibco BRL) in a 5 % CO, incubator at 37°C. Cells (~8 x 105) were plated on 100-mm diameter tissue culture dishes (Corning, NY) and cultured for ~48 hr before addition of 66 treatments. C.2. Materials Fumonisin B, used in this study was purchased from PROMEC (South Africa) and actinomycin D (4 ng/ml) was obtained from Sigma (St. Louis, MO). Sphinganine and sphinganine l-phosphate were purchased from Biomol (Plymouth Meeting, PA). Sphinganine was prepared by forming a 1 :1 complex with Bovine Serum Albumin (cell culture grade, GibcoBRL). Sphinganine l-phosphate was dissolved in methanolztetrahydrofirranzwater (60:30: 1 0) at 65 °C, evaporated, and dissolved in 0.4% BSA solution. RNA Stat-60 was from Tel-Test, Inc (Friendswood, TX). C.3. Analysis of DNA fragmentation by agarose gel electrophoresis assay mg treatments, fragmented DNA from LLC-PK] cells was extracted as described (Sellins and Cohen, 1987). In brief, cells from IOO-mm dish were harvested using rubber policeman and collected by centrifugation (5 min, 500 g). The medium was removed and the cell pellet was resuspended in phosphate-buffered saline (PBS). After another centrifuge, the pellet was suspended in 0.1 m1 hypotonic lysis buffer (10 mM Tris, 10 mM EDTA, 0.5% Triton X—100, pH 8.0). Cells were incubated for 10 min at 4°C and the lysate was centrifuged (30 min, 13,000 g, 4 ° C). The supernatant containing fragmented DNA was treated with RNase A (0.4 pg/pl) for 1 hr at 37°C and then incubated with proteinase K (0.4 [lg/pl) for 1 h at 37°C. DNA was precipitated overnight by incubation at -20°C in 50% isopropanol and 0.5 M NaCl. The DNA precipitate was pelleted by centrifugation (30 min, 13,000 g, 4°C), dissolved in distilled H,O and then electrophoresed in a 2 % agarose gel containing ethidium bromide (0.5 [lg/ml) in 0.5X Tris-borate-EDTA buffer. After electrophoresis, DNA was visualized with a UV transilluminator and photographs were taken using a Polaroid system. C.4. Differential display reverse transcription polymerase chain reaction (DDRT- 67 PCR) Differential display gel kits including RTase (Genhunter Corp., Nashville, TN) were used for this study as described previously (Kim et al., 2001). CS. Immunoblot analysis for calmodulin Cellular protein was extracted from LLC-PK, cells with lysis buffer (20 mM Tris-Cl, pH 8.0, 150 mM NaCl, 10 % glycerol, 1 % Triton X-100, 2 mM EDTA, 1 mM penylmethylsulfonyl fluoride, 10 BM of leupeptin and aprotinin cocktail solution), sonicated, and centrifuged. The supernatant protein was denatured by incubating for 5 min at 95 °C in Laemmli buffer. The sample was applied to 15 % SDS-bis acrylamide denaturing gel and transferred onto nitrocellulose membrane. The protein on the membrane was blocked by overnight incubation in 5 % non-fat dry milk/TBS- Tween 20 (0.1%), pH 7.4. The membrane was probed with anti-calmoduin antibody (Santa Cruz) raised against a goat calmodulin for 1 hour at room temperature and treated with anti- goat IgG conjugated with horseradish peroxidase for additional 1 hour at room temperature. Chemiluminescence solution from Santa Cruz was used to develop the immunoblot. C.6. Northern blot analysis Total cellular RNA was extracted from LLC-PK, cells by the modification of the one-step method (Chomczynski and Sacchi, 1987) using RNA Stat-60 reagent (Tel-test “B” Inc., Friendswood, TX). 20 pg of total RNA was separated by electrophoresis on a 1 .2% (wt/vol) agarose gel (Gibco BRL) containing 0.6 M formaldehyde (Sigma), 0.02 M MOPS (Gibco BRL) at 80 Volt for ~3 hr. The gel was stained with ethidium bromide to verify the integrity and equal RNA loading, and the RNA was transferred to a nylon membrane (NEN, Boston, MA) with 20X SSC. PCR-amplified calmodulin cDNAs (Kim et al. , 2001) labeled with 50 pCi a-32P-dCTP (DuPont NEN) were prepared by random oligo-labeling method (Prime-a-Gene Labeling System, Promega). Afier cross-linking, the membrane was pre-hybridized at 42°C in Denhardt’s hybridization 68 buffer (Kingston, 1997) for 2 hr. Blots were hybridized at the same temperature overnight in a hybridization bag containing radio-labeled cDNA probe described earlier. The membrane was then washed in low and high stringency washing buffer and exposed to the K-screen (Kodak, Rochester, NY) for phospho-imager analysis. Density of each band was quantified using the FX software program after background was subtracted. C.7. Statistical Analysis The statistical analysis was performed by using Sigma-Stat Analysis system (Jandel Scientific, San Rafael, CA). The half lives for the calmodulin mRNAs were calculated using regression analysis. The student t-test was used to analyze calmodulin mRNA for statistical significance. Differences were considered statistically significant at p .<_ 0.05. D. RESULTS LLC-PK, renal kidney epithelial cells were used as a model system for these studies because the kidney is the most sensitive organ to fumonisin toxicity (Bucci et al., 1998; Gumprecht et al., 1995). Furthermore, fumonisin B, induced expression of calmodulin in LLC-PK, cells and apoptosis induced by the mycotoxin was dependent on calmodulin (Kim et al., 2001). F umonisin B, increases calmodul in mRNA and calmodulinpratein in concentration- dependent manners. Fumonisin B, induces calmodulin protein in a concentration- dependent manner (Kim et al. , 2001). To compare the response of calmodulin mRNA to that of calmodulin protein, kidney cells were cultured with various concentrations of fumonisin B,. Fumonisin B, induced both calmodulin mRNA (Figure 3.2A) and calmodulin protein (Figure 3.28) in concentration-dependent manners. At 25 pM, fumonisin B, induced 69 A CaM mRNA9 ! fl 8 a 700 e 600 E 8 500 .5, 400 g 300 200 E 100 E, 0 0 . 5 50 Fumonrszrn B, (11M) B ,3 250 g 200 U 8\: 150 .E 100 8 e 50 ad 0 E: o _25 50 U Fumomsrn B, (PM) Figure 3. 2. Fumonisin B, increases calmodulin mRNA and calmodulin protein in ‘ ..tL...‘ ‘ manners. 106 LLC- PK, kidney cells were plated 1n 100 mm dishes and grown for 48 hrs. Cells were then added with fumonisin B, at 0, 10, 25, and 50 11M 1n 3 fresh culture medium and cultured for 16 hr. At the end of the culture period, RNA (A) and cellular protein (B) were harvested and Northern and Western analysis were performed respectively as described under “Materials and Methods”. 70 calmodulin mRNA ~3-fold, while increasing calmodulin protein ~70%. At 50 uM, fumonisin B, increased calmodulin mRNA ~5-fold and caused a corresponding increase in calmodulin protein of ~2.4 fold. F umanisin B, does not affect stability of calmodulin mRNA. I n c r e a s e d abundance of calmodulin mRNA by fumonisin B, could be either due to increased gene transcription or increased mRNA stability. To examine the effect of the mycotoxin on stability of calmodulin mRNA, the transcription inhibitor actinomycin D was used to prevent transcription. In the absence of the mycotoxin, the concentration of calmodulin mRNA was initially increased up to 6 hr and then decreased steadily until 36 hr (Figure 3.3). Regression analysis applied to the data from 6 to 36 hour indicates that calmodulin mRN A has a half life of ~30 hr. A similar pattern of calmodulin mRNA was observed in response to fumonisin B, with an mRN A half life of ~36 hr; however, regression analysis indicates that the slopes are not significantly different between the two groups, indicating that fumonisin B, does not affect the stability of calmodulin mRN A. Novel gene transcription is required for fumonisin B ,-induced apoptosis. Since fumonisin B, appears to increase calmodulin expression by increasing transcription rather than by increasing mRN A stability and apoptosis is dependent on calmodulin, dependence of fumonisin-induced apoptosis on novel gene transcription was examined using actinomycin D. Fumonisin B, alone at 50 BM caused DNA fragmentation (Figure 3.4); whereas, actinomycin D at 2 ng/ml has no effect. However, addition of actinomycin D in the presence of fumonisin B, prevented fumonisin-induced DNA fragmentation. F umanisin B, selectively induces calmodulin. To determine whether fumonisin B, induces calmodulin in a selective manner, DDRT-PCR was conducted as previously 71 0 6 --....r Wax... '4‘ -.,,. vb- : ~ - . _“‘“".W V hi 1 I ' ‘ ’ ‘ ‘ i . "f" V'mmgt" L... _ 1:: “a. . ,1 . ‘ l ‘7 - . - ‘ _ - _ . ,_ > , . .. . _ y 7”” :7" “’7‘ ’5" \W- “a 3' ‘21:" """ I‘ " , .‘n L . .. _, r.‘ " . , I, r .. 2 ’ ’ . . 1’ . . ~ . . ‘ r. ' " ‘1 " . . . . , ‘1 "-'-"'W-. t 5* We’mJYIC- ‘ - an". “I. t‘ I 300000 ‘ 250000 150000 100000 ~ +Com, | 50000 ,. "O—Fumonrsm B1 l I 1 1 1 1 0 510.15 20 25 30 35 40 Hours CaM mRNA (CNT*mm2) Figure 3.3. Fumonisin B, does not affect stability of calmodulin mRNA. 10‘5 LLC-PK, kidney cells were plated in 100 mm dishes and grown for 48 hrs. Actinomycin D (4 ng/ml) was then added to culture media in the presence or absence of fumonisin B, and kidney cells were cultured at various hours. At the end of the culture, total RNA was harvested and Northern analysis was performed as described under “Materials and Methods”. 72 ActD — + + (2 ng/mL) FB, + - + (5011M) Figure 3.4. Novel gene transcription is required for fumonisin B,-induced apoptosis 10" LLC-PK, kidney cells were plated in 100 mm dishes and grown for 48 hrs. Cells were then added with 50 11M fumonisin B, (F B,) and/or 2 ng/ml actinomycin D (ActD) in a fresh culture medium and cultured for ~48 hr. At the end of the culture, fragmented DNA was harvested and analyzed by agarose gel electrophoresis as described under “Materials and Methods”. 73 described (Kim et al., 2001). The results demonstrate that an apoptotic concentration of fumonisin B, affected expression of only a few genes in LLC-PK, cells (Figure 3.5). Those genes include calmodulin (CaM). Depletion of complex sphingolipids does not affect expression of calmodulin. Since fumonisin-induced kidney cell apoptosis requires calmodulin activity (Kim et al., 2001) and appears to be partially dependent on depletion of complex sphingolipids (Yoo et al., 1996; Riley et al., 1999), the effect of depletion of complex sphingolipids on calmodulin mRN A was examined by using B-fluoroalanine. B-fluoroalanine blocks de novo sphingolipid biosynthesis similar to fumonisin B, but without causing accumulation ofsphinganine (3.1). The concentration of calmodulin mRNA was not altered by B-fluoroalanine (Figure 3.6), suggesting that depletion of complex sphingolipids does not play a role in fumonisin-induced calmodulin expression. Sphinganine does not mediate fumonisin-induced calmodulin expression. [3- fluoroalanine blocks sphinganine accumulation caused by fumonisin B, and partially prevents calmodulin expression (Kim et al., 2001). To determine whether sphinganine acts directly to induce calmodulin expression, exogenous sphinganine at 10 pras added to LLC-PK, cultures. As shown in Figure 3.7, addition of an apoptotic concentration of exogenous sphinganine had no effect on calmodulin mRNA. Sphinganine I-phosphate induces calmodulin expression. Since sphinganine 1- phosphate accumulates in the presence of fumonisin B, (Smith and Merrill, 1995), the effect of this sphinganine metabolite on calmodulin expression was also examined. Sphinganine l-phosphate at 2 11M increased the concentration of calmodulin mRNA ~50% in LLC-PK, kidney cells (Figure 3.8). 74 Figure 3.5. Fumonisin B, selectively induces calmodulin. Cells were cultured in the absence (C) or presence (F) of 50 11M fumonisin B, for 16 hr and DD RT-PCR was performed as described under “Materials and Methods”. Arrows indicate differential expression caused by fumonisin B, (CaM=calmoduIin). The primers used for this figure were 5'-T12-A-3' and 5'-TAGCAAGTGC-3'. 75 Control BFA W CaMmRNA-) ' “ * * 16000 ‘ “‘"" """'” 14000 12000 10000 8000 6000 4000 2000 0 CaM mRNA (CNT *mm2) Control BFA Treatment Figure 3.6. Depletion of complex sphingolipids does not affect expression of calmoduolin. LLC-PKI cells were cultured in the presence of 50 |J.M B-fluoroalanine (BF A) for 16 hr. At the end of 16 hr culture, total RNA was harvested and Northern analysis was performed as described under “Materials and Methods”. 76 Control m Sphinganine CaM mRNA-) 1 6000 14000 1 2000 l 0000 8000 6000 4000 2000 0 CaM mRNA (CNT *mm2) Control Sphinganine Treatment Figure 3.7. Sphingoid bases does not mediate fumonisin-induced calmodulin expression. LLC-PK, cells were cultured in the presence of 10 pM sphinganine. At the end of culture, total RNA was harvested and Northern analysis was performed as described under “Materials and Methods”. 77 Control CaM mRNA->145 ‘ 2000 1 800 1 600 1400 1200 1 000 800 600 400 200 0 CaM mRNA (CNT*mm2) Control SA— 1 -P Treatment Figure 3.8. Sphinganine l-phosphate induces calmodulin expression. LLC-PK, cells were cultured in the presence of 2 1.1M sphinganine l-phosphate(SA-1-P) for 3 hr. At the end of culture, total RNA was harvested and Northern analysis was performed as described under “Materials and Methods”. 78 E. DISCUSSION The mechanism of toxicity of fumonisins has not been well elucidated, but an apoptotic concentration of fumonisin B, was shown to induce calmodulin expression in LLC- PK, renal tubular epithelial cells (Kim et al., 2001). Moreover, fumonisin-induced kidney cell apoptosis was dependent on calmodulin activity (Kim et al., 2001). Calmodulin is the major Ca” sensor protein and plays an essential role in regulating many cellular processes including cell proliferation (Rasmussen and Means, 1987; Rasmussen and Means, 1989), cell-cycle progression (Chafouleas et al., 1982; Chafouleas et al., 1984; Rasmussen et al., 1992), transformation (Chafouleas et al., 1981) and apoptosis (Dowd et al., 1991; Kim et al., 2001; Rosenthal et al., 1998). Calmodulin is ubiquitous and structurally conserved across species and regulates more than 30 enzymes in mammalian cells via the Caiilcalmodulin complex (Stoclet et al., 1987). The activity of calmodulin was long thought to be regulated primarily by changes in the net flux or distribution of Ca” rather than by changes in the cellular concentration of the protein because expression of calmodulin is tightly regulated (Chafouleas et al., 1981 ). Results of the present study show that fumonisin B, induces both calmodulin mRNA and calmodulin protein in LLC-PK, kidney cells in concentration- dependent manners. Furthermore, DDRT-PCR clearly demonstrates that fumonisin B, causes differential expression of calmodulin and only a few other genes. This is consistent with our previous findings indicating specificity in response to fumonisin B, (Kim et al., 2001) and with studies showing early induction of calmodulin gene expression in lymphocytes undergoing glucocorticoid-mediated apoptosis (Dowd et a1 ., 1991). In addition, changes in the concentration of calmodulin have been reported in differentiated BC3H-1 mouse myoblast cells (Epstein et al., 1989), in NGF-induced differentiation of neuronal cells 79 (Bai and Weiss, 1991), in CHO-K1 cells released from stationary phase to re-enter into the cell cycle (Chafouleas et al., 1984), in transformed chicken embryo fibroblasts (Watterson et al., 1976), and in several malignancies including those of liver (Wei et al., 1982), lung (Liu et al., 1996) and breast (Singer et al., 1976; Tuccari et al., 1993). Increased concentrations of calmodulin caused by fumonisin B, may stimulate a variety of calmodulin- dependent enzymes (Heist and Schulman, 1998; Van Eldik et al, 1979) which could, in turn, mediate the kidney toxicity of fumonisins by causing renal tubular cell apoptosis. Increases in calmodulin mRNA caused by fumonisin B, could be due to either increased transcription of the gene or increased stability of the mRNA. The present study supports a role for fumonisin B, in increasing calmodulin transcription since fumonisin B, did not alter the stability of calmodulin mRNA. The half life of calmodulin mRNA in LLC- PK, cells is ~30 hr and is not significantly altered by fumonisin B,. Bai and Weiss (1991) have reported comparable calmodulin mRNA half lives in PC 12 neuronal cells ranging from ~10 to 29 hr depending on the size of the transcripts. This is consistent with the results of other studies which suggest that the steady state level of cellular calmodulin is increased by a higher rate of transcription rather than changes in mRNA stability (Chafouleas et al., 1981 ; Epstein et al., 1989). The mechanism by which fumonisin B, increases calmodulin transcription may involve disruption of sphingolipids. Numerous studies both in viva and using a variety of cell types have demonstrated that the mechanism of toxicity of fumonisins requires disruption of sphingolipid metabolism via inhibition of sphinganine N- acyltransferase (Merrill et al., 1997a). There are several possible mechanisms by which disruption of sphingolipid metabolism could cause induction of calmodulin. First, fumonisin B, could increase 80 calmodulin expression by depleting complex sphingolipids. Several studies have demonstrated that depletion of complex sphingolipids mediate, at least in part, fumonisin- induced toxicities (Yoo et al. , 1996; Riley et al. , 1999). When B-chloroalanine was used to inhibit serine palmitoyltransferase and block de nova synthesis of sphingolipids in LLC-PK, cells to a similar degree as fumonisin B,, it both inhibited cell growth and increased cell death (Yoo et al., 1996). Also, the addition of ceramide derivative reversed fumonisin inhibition of axonal growth in hippocampal neurons (Harel and F uterman, 1993) and the addition of N—acetylsphinganine partially protected human keratinocytes from fumonisin B,- induced apoptosis (Tolleson et al., 1999). Riley et al (1999) demonstrated that prevention of fumonisin-induced sphinganine accumulation by inhibitors of serine palmitoyltransferase blocks fumonisin-induced apoptosis but does not abolish the morphological changes caused by the mycotoxin, suggesting that depletion of sphingolipids could play a role in fumonisin toxicity. To examine the role of depletion of complex sphingolipids in fumonisin—induced calmodulin induction, B-fluoroalanine was used to block de novo sphingolipid biosynthesis. However, the findings demonstrate no effect of B-fluoroalanine on calmodulin mRNA and, therefore, suggest that depletion of complex sphingolipids does not mediate fumonisin- induced calmodulin expression in LLC-PK, kidney cells. A second possible mechanism whereby disruption of sphingolipid metabolism could mediate fumonisin-induced calmodulin expression is by causing accumulation of the sphingoid base sphinganine. B-fluoroalanine together with fumonisin B, blocks fumonisin- induced sphinganine accumulation and partially prevents fumonisin-induced calmodulin expression in LLC-PK, kidney cells (Kim et al., 2001). However, the addition of exogenous sphinganine did not affect calmodulin mRNA. Therefore, sphinganine does not appear to 81 directly mediate fumonisin-induced calmodulin expression. Another possible mechanism by which disruption of sphingolipid metabolism could mediate fumonisin-induced calmodulin expression is via the sphinganine metabolite, sphinganine l-phosphate. Smith and Merrill (1995) showed that in the presence of fumonisin B,, sphinganine is converted to the l-phosphate metabolite by sphingosine (sphinganine) kinase. The present study demonstrates that exogenous sphinganine 1- phosphate induces calmodulin expression and suggests that increased formation of sphinganine l-phosphate in the presence of fumonisin B, in kidney cells, at least in part, mediates stimulation of calmodulin expression. Studies have demonstrated that intracellularly generated sphingoid base l-phosphate can be released into extracellular space in a variety of cells (Goetzl and An, 1998; Hla, 2001; Spiegel and Merrill, 1996). Therefore, it is possible that both endogenously produced sphinganine l-phosphate as a result of fumonisin B, exposure as well as exogenously added sphinganine l-phosphate may act through interaction with G-protein coupled sphingoid base l-phosphate receptor EDG family members to induce calmodulin expression. The lower effect of exogenous sphinganine 1- phosphate on calmodulin expression than fumonisin B, may be due to complete degradation of exogenously added sphinganine l-phosphate to hexadecanal and ethanolamine phosphate (Merrill et al., 1996); whereas, in the presence of fumonisin B,, endogenously generated sphinganine l-phosphate would continue to be produced. The mechanism by which sphinganine l-phosphate increases calmodulin expression is not known, but may involve mobilization of intracellular Ca“. Changes in intracellular Ca’+ have been shown to affect changes in nuclear functions including gene expression (Bito et al., 1997). Sphingoid base l-phosphate has been shown to increase cytosolic Ca++ by 82 mobilizing Ca” from internal stores (Ghosh et al., 1994; Mattie et al., 1994; Tomquist et al., 1997). Elevated cytosolic Ca” is accompanied by significant elevation of nuclear Ca”, producing a signal which may directly activate Caii-dependent proteins within the nucleus (Heist and Schulman, 1998; Himpens et al., 1994). However, Ca” mobilization by sphinganine l-phosphate is not likely to be the mechanism of calmodulin induction by fumonisin B, because addition of intracellular Ca“ chelator BAPTA AM did not affect fumonisin-induced calmodulin increase (data not shown). Another possible mechanism by which sphinganine l-phosphate might induce calmodulin is via activation of the MAPK pathway and transcription factor Sp-l. Toutenhoofd et al (1998) demonstrated that human calmodulin is regulated at the transcriptional level, that the 5' untranslated region of the calmodulin gene was necessary, and that it contains multiple Sp-l sites (Toutenhoofd et al., 1998). The Sp transcription factor family is necessary for basal calmodulin gene transcription and for regulation of calmodulin expression (Pan et al., 2000; Solomon et al., 1997). Furthermore, Zhang et al (1999) demonstrated that fumonisin B, activates transcription of CDK inhibitor p21 via Sp- lbinding sites. It is not clear if sphingoid base l-phosphate activate Sp—l binding to DNA; however, the l-phosphate has been shown to activate mitogen-activated protein kinase (MAPK) which, in turn, activates a variety of transcription factors including Sp-l (Davis et al., 1996). Therefore, it is plausible that fumonisin B, causes elevated sphinganine 1- phosphate which, in turn, activates the MAPK pathway and transcription factors such as Sp- 1, resulting in stimulation of calmodulin transcription. In the cell, the Caii/calmodulin complex plays a wide variety of roles and several studies reported evidence relating calmodulin to apoptosis, especially those induced by 83 sustained increases in cellular calcium concentration (Kim et al. , 2001; Bellomo et al., 1992; Dowd et al., 1991). Inhibition of calmodulin activity by W7 protects LLC-PK, cells fi'om fumonisin-induced apoptosis (Kim et al. , 2001). In addition, treatment of certain lymphoma cell lines and thymocytes with glucocorticoid activates a programmed cell death pathway (McConkey et al., 1989) and calmodulin inhibitors also protected lymphocytes from glucocorticoid-induced apoptosis (Dowd et al., 1991). Also, calmodulin antagonists inhibited F as-mediated apoptosis of CD4+ T-cells from patients with AIDS (Pan et al., 1998) and sulfur mustard-induced apoptosis of keratinocytes (Rosenthal et al., 1998). Despite these findings, the mechanism by which calmodulin is involved in the signaling of the cell-death pathway is not understood. A recent study discovered a new calcium/calmodulin-dependent protein kinase called death-associated protein kinase 2 that is believed to signal apoptosis through its catalytic activity (Kawai et al., 1999). Further studies are needed to determine if this novel protein kinase is involved in fumonisin-induced kidney cell apoptosis. Taken together, the findings of the present study provide a possible mechanism by which fumonisin B, induces calmodulin expression in LLC-PK, porcine renal tubular epithelial cells. Fumonisin B, disrupts sphingolipid metabolism and causes accumulation of sphinganine l-phosphate which activates transcription of calmodulin. Further studies are necessary to elucidate the underlying mechanism of calmodulin induction by sphinganine 1- phosphate and the role of increased concentration of calmodulin in fumonisin-induced toxicity. 84 CHAPTER IV FUMONISIN B, AND SPHINGANINE REDUCE PHOSPHORYLATION OF MAPK/ERK AND PKB/AKT IN LLC-PK, PORCINE RENAL TUBULAR EPITHELIAL CELLS 85 A. ABSTRACT Fumonisins are a family of mycotoxins produced by F usarium moniliforme which is the most common mold found on corn throughout the world. These compounds are both toxic and carcinogenic for animals, and perhaps humans, with the kidney being the most sensitive organ to firmonisins. F umonisin B,, the most prevalent of these mycotoxins, induces apoptosis in LLC-PK, renal tubular epithelial cells (Kim et al., 2001). The molecular mechanism of fumonisin toxicity is not clear, but appears to involve disruption of de nova biosynthesis of sphingolipids. Fumonisin B, inhibits ceramide synthase causing depletion of complex sphingolipids and accumulation of the bioactive sphingoid base, sphinganine. In the present study, LLC-PK, cells were used to examine the effects of disruption of sphingolipid metabolism on signaling pathways responsible for determining cell survival and cell death. The studies used B-fluoroalanine as a tool to dissect the relative contribution of depletion of complex sphingolipids and accumulation of sphinganine. B- fluoroalanine is an inhibitor of serine palmitoyltransferase which, when added alone, blocks complex sphingolipid biosynthesis to a similar degree as firmonisin B, , but without causing accumulation of sphinganine. When added together with fumonisin B,, B-fluoroalanine blocks sphinganine accumulation. F umonisin B, alone had no apparent effect on kidney cell growth or death within the first 16 hrs of culture, but inhibited kidney cell growth and caused cell death with 24-36 hrs of culture, consistent with dependence on disruption of sphingolipid metabolism. Similarly, the exogenous addition of sphinganine inhibited growth and caused death, but within only 1-2 hrs of culture. F umonisin B, had no effect on ERK phosphorylation at 2 hrs of culture, but reduced phosphorylation of MEK and ERK at 16 hrs. In contrast, B-fluoroalanine did not decrease phosphorylation of MEK or ERK together with 86 err-‘91 fumonisin B,. Similar to fumonisin B,, sphinganine alone and in the presence of a sub- optimal concentration of fumonisin B, reduced phosphorylation of both MEK and ERK. Neither fumonisin B,, B-fluoroalanine, nor sphinganine affected either total ERK or phosphorylation of Raf-1. Both fumonisin B, and sphinganine decreased phosphorylation of Akt at both serine 473 and threonine 308; whereas, B-fluoroalanine did not decrease phosphorylation of Akt and blocked the effect of fumonisin B,. The findings indicate that growth arrest and apoptosis induced by fumonisin B, in LLC-PK, porcine renal tubular epithelial cells is mediated by sphinganine which inhibits phosphorylation of Akt and of ERK by a Raf-l independent pathway. B. INTRODUCTION Fumonisins are cytotoxic and carcinogenic mycotoxins produced by F usarium moniliforme, the most common mold contaminant of corn and other agricultural commodities in the United States and throughout the world (Marasas, 1996). Fumonisin B,, the major fumonisin, causes several animal diseases including equine leukoencephalomalacia (Marasas et al. , 1988a) and porcine pulmonary edema (Osweiler et al. , 1992). Early studies with impure culture material containing fumonisins indicated that the mycotoxins cause liver tumors (Gelderblom et al., 1996a; Gelderblom et al., 1996b). The National Toxicology Program has confirmed the carcinogenicity of fumonisin B, with a two year study which demonstrated that fumonisin B, induces hepatic tumors in female B6C3F mice and renal tubule tumors in male F 344 rats (Howard et al., 2001). In addition, consumption of F usarium-contaminated corn has been associated with human esophageal cancer in certain areas of the world (Marasas et al. , 1988b). F umonisin B, is wide spread in variable amounts 87 among food-grade corn and com-based food products of US. origin (Gutema et al., 2000) suggesting that human populations are at significant risk of exposure to fumonisin B, through their normal diet. This raises concerns regarding possible health implications that can be caused by chronic exposure to these mycotoxins. The toxicity caused by fumonisin B, is mediated by disruption of sphingolipid metabolism. Due to its structural similarity with sphingolipids (Figure 4.1), fumonisin B, potently and competitively (with both sphinganine and acyl CoA) inhibits sphinganine (sphingosine) N-acyltransferase (ceramide synthase), the enzyme that acylates sphinganine to form ceramide in the de novo biosynthetic pathway of sphingolipids (Merrill et al. , 1993a; Wang et al. , 1991). Inhibition of ceramide synthase causes accumulation of sphinganine and simultaneous depletion of complex sphingolipids including sphingomyelin as shown in mice (Martinova and Merrill, 1995), rats (Kwon et al., 1997; Riley et al., 1994), mink (Restum et al., 1995), ponies (Wang et al. , 1992) and pigs (Riley et al., 1993a). Since sphingolipids are bioactive compounds that regulate cellular behavior, i.e., cell proliferation, terminal differentiation, and apoptosis (Hannun and Bell, 1989; Spiegel et al., 1994; Zhang et al., 1991), firmonisin-induced disturbance of the balance among these lipid molecules has significant impact on the fate of cells, leading to degenerative diseases. Sphingoid bases like sphinganine are growth-inhibitory and apoptotic (Jarvis et al., 1996; Ohta et al., 1994, 1995) and accumulation of sphinganine and its l-phosphate metabolite appear, at least in part, to be responsible for fumonisin-induced toxicity (Kim et a1. , 2001; Kim and Schroeder, 2002). F umonisin B, induces apoptosis in LLC-PK, porcine renal tubular epithelial cells in a sphinganine-dependent manner (Kim et al. , 2001). The growth inhibitory and apoptotic effects of fumonisin B, could be mediated by 88 0% CH,COO(-) \\c- -CH,CHCOO(-) Fumonisin B, 0 OH OH CH3 CH3 (I) CH3 0H NH, C-CH,CHCOO(-) /, l O CH,COO(-) OH OH Sphinganine AA/VWVW NH, OH OH Sphingosine WW NH, Figure 4.1. Structures of fumonisin B,, sphinganine, and sphingosine. 89 sphinganine via inhibition of the mitogen-activated protein kinase (ERK) pathway, a crucial regulator of growth, survival, and maintenance of various kidney cells (di Mari et al., 1999; Dudley et al., 1995; Ishizuka et al., 1999; Kinane et al., 1997; Schramek et al., 1997). Fumonisin B, (Huang et al., 1995) and sphingoid bases (Hannun and Bell, 1986; Smith et al., 2000) potently inhibit protein kinase C (PKC), an important activator of MEK and p44/p42 MAPK (or ERK), via either Raf-1 or MEK kinase. Inhibition of ERK activity by sphingoid bases has been suggested to be a partial mechanism of action in sphingoid base- induced growth arrest and apoptosis in various tumor cell lines (Sakakura et al., 1998; Jarvis et al. , 1997). F umonisin-induced sphinganine accumulation may suppress ERK activity and contribute to fumonisin-induced growth inhibition and cell death via inhibition of PKC. Alternatively, the growth inhibitory and apoptotic effects of fumonisin B, may be mediated by sphinganine via suppression of P13 K/Akt signaling, an important anti-apoptotic pathway (Krasilnikov, 2000; Zhou et al., 2000). PI3K/Akt signaling has been shown to mediate both Ras- and cytokine-induced protection from apoptosis. Activation of PI3K phosphorylates Akt (Stephens et al., 1998; Stokoe et al., 1997), a serine/threonine kinase which, in turn, suppresses apoptosis by phosphorylating Bad and caspase-9 (Krasilnikov, 2000; Zhou et al., 2000; Zundel and Giaccia, 1998). The sphingoid base sphingosine has been shown to reduce phosphorylated Akt with concomitant reduction of Akt activity during sphingosine-induced apoptosis in human hepatoma cells (Chang et al., 2001). Also, constitutively active Akt prevents sphingosine-induced apoptosis (Chang et al., 2001). Thus, sphinganine accumulation caused by fumonisin B, may suppress Akt phosphorylation and trigger a molecular cascade leading to growth inhibition and apoptosis. The present study used LLC-PK, renal tubular epithelial cells to examine the effects of 90 disruption of sphingolipid metabolism on ERK and PI3K/Akt phosphorylation cascades leading to LLC-PK, kidney cell growth arrest and apoptosis. To determine the relative contributions of depletion of complex sphingolipids and accumulation of sphinganine on ERK and Akt phosphorylation, B-fluoroalanine, an inhibitor of an earlier step in de nova sphingolipid biosynthesis than fumonisin B,, was used both alone and together with fumonisin B,. C. MATERIALS AND METHODS C]. Materials. Fumonisin B, was purchased from Sigma and dissolved in phosphate buffered saline (PBS). D-erythra-sphinganine was purchased from Biomol Research Laboratories, Inc. and prepared for cell culture treatment as previously described (Schroeder et al., 1994). Rabbit antibodies against ERK, pERK, Akt, and pAkt proteins were obtained from New England BioLabs (Beverly, MA) and anti-rabbit IgG was obtained from Santa Cruz. C.2. Cell Culture. Porcine kidney tubular epithelial LLC-PK, cells were obtained from American Type Culture Collection and cultured in Dulbecco’s Modified Eagle Medium/F- 12K nutrient solution (Gibco BRL) (1:1) with 5% fetal bovine serum (F BS) and 100 pM L- serine (Gibco BRL) until they reach the appropriate confluency for chemical treatments in each experiment. C.3. Total Nucleic Acid Measurement. Briefly, 105 cells were cultured in 6-well plates for 24 hr and treated for appropriate times. Then, cells were washed with ice-cold PBS twice and air-dried. Cellular nucleic acid was harvested by lysing cells with 0.1 N NaOH and quantitated by measuring absorbance at 260 nm. 91 C.4. Western blot Analysis. After treatment as indicated in the figure legend, cells were washed in ice-cold PBS once and cellular protein was extracted from LLC-PK, cells with lysis buffer (20 mM Tris-Cl, pH 8.0, 150 mM NaCl, 10% glycerol, 1% Triton X-100, 2 mM EDTA, 1 mM phenylmethylsulfonyl fluoride, 10 11M of leupeptin and aprotinin cocktail solution) and sonicated for 15-20 sec followed by centrifuge. The protein extract was denatured by incubating for 5 min at 95°C in Laemmli buffer. The sample was applied to SDS-bisacrylamide denaturing gel and transferred onto PVDF membrane (NEN Life Sciences). The proteins on the membrane were blocked for an hour in 5% non-fat dry milk/TBS-0.1% Tween 20 solution. The membrane was incubated in 3% BSA/TBS-T containing primary antibody by overnight followed by 1 hr incubation in blocking solution containing anti-rabbit IgG conjugated with horseradish peroxidase and developed with chemiluminescence solution from Santa Cruz. C.5. Statistical Analysis. The statistical analysis was performed by using Sigma-Stat Analysis system (Jandel Scientific, San Rafael, CA). The student t-test was used to analyze data for total nucleic acids for statistical significance. Differences were considered statistically significant at p s 0.05. D. RESULTS F umonisin B, inhibits LLC-PK, cell growth and induces apoptosis. To investigate the effect of fumonisin B, on proliferation, LLC-PK, cells were cultured in the absence or presence of 50 |J.M fumonisin B, for 48 hr and total nucleic acids were measured as an index of cell number (Figure 4.2). In control cultures, total nucleic acids increased at a consistent rate throughout the entire 48 hr culture period with a doubling time of ~1 8 hr. Fumonisin 92 on O NWAUIO‘Q OOOOOO Total Nucleic Acids (pg/well) S O 6 12 18 24 30 36 42 48 Hours 0 Figure 4.2. Fumonisin B, inhibits growth and induces death of LLC-PK, kidney cells. ~105 LLC-PK, cells per well were cultured in a six-well plate in the absence (E1) or presence of 50 11M fumonisin B, (I) at various times up to 48 hrs. At each time point, cells were harvested and total nucleic acids were measured as described under “Materials and Methods”. * indicates statistical difference between two groups (P<0.0001). :20 T) E 315- U) 2 310_ . U 1., 3 5* 2 ~55 “a 60 1 1 1 A 1 ‘- 0 2 6 8 10 Figure 4.3. Sphinganine inhibits growth and induces death of LLC-PK, kidney cells. ~105 LLC-PK, cells per well were cultured in a six-well plate in the absence (0) or presence of 20 11M D-erythro-sphinganine (O) at various times. At each time point, cells were harvested and total nucleic acids were measured as described under “Materials and Methods”. * indicates statistical difference between two groups (P<0.0001). 93 B, produced morphological changes characteristic of apoptosis within 16-24 hr, including chromatin condensation, cell shrinkage, membrane blebbing, and formation of apoptotic bodies (data not shown). Total nucleic acids in cells cultured with fumonisin B, increased at a similar rate to that of control cultures up to 24 hr. Thereafter, total nucleic acids did not change during the final 24 hr of culture. Sphinganine inhibits LLC—PK, cell growth and induces apoptosis. To determine if sphinganine causes similar effects on kidney cell growth to that of fumonisin B,, LLC-PK, cells were cultured in the presence of 20 1.1M sphinganine and total nucleic acids were measured as an index of cell number. In the presence of sphinganine, LLC-PK, cells showed morphological changes characteristic of apoptosis similar to those of the cells cultured with fumonisin B,, but as early as 1 hr after addition of the sphingoid base (data not shown). Simultaneously, sphinganine significantly decreased total nucleic acids within 1-2 hrs and killed 70% of the cells within 4 hr (Figure 4.3). F umonisin B, inhibits ERK and MEK phosphorylation in a sphinganine-dependent manner. Fumonisin B, does not cause significant sphinganine accumulation in LLC- PK, cells until 6 hr of culture (Kim et al., 2001). To determine the effect of fumonisin B, on ERK phosphorylation prior to sphinganine accumulation, LLC-PK, kidney cells were cultured in the presence of 50 |.1.M fumonisin B, for various times up to 2 hr and cellular protein was lysed and used for Western blot analysis with an antibody specific for phosphorylated ERK @ERK). Fumonisin B, increased pERK rapidly with maximum phosphorylation at 10 min. Thereafter, phosphorylation decreased to the basal level within 30 min (Figure 4.4A). To determine if the effect was specifically caused by fumonisin B,, pERK of cells cultured with fumonisin B, was compared to that of control cultures. Time- 94 O 5 10 30 60 120min Figure 4.4. Time-dependent changes in the phosphorylation of ERK in control and fumonisin B,-treated LLC-PK, cells. LLC-PK, cells were grown in DMEM/F12-K growth medium containing 5% Fumonisin B, for 48 hours in 100 mm dishes. A. Kidney cells were treated with 50 BM of fumonisin B, for 5, 10, 30, 60 and 120 minutes and harvested for western blot analysis against phosphorylated form of ERK (pERK) protein as described under “Materials and Methods”. B. Comparison between control (C) and fumonisin B, (F) at 0, 10, and 30 min. 95 dependent changes in phosphorylation of ERK (pERK) were similar for both control cultures and fumonisin-treated cultures (Figure 4.48). To further examine the effect of fumonisin B, on ERK phosphorylation and the role of sphinganine accumulation, additional studies were conducted for 16 hr. This is the point at which cellular sphinganine increases to ~1 nmol/mg protein and DNA fragmentation is evident (Kim et al. , 2001). B-fluoroalanine, which inhibits the first step of de nova sphingolipid biosynthesis catalyzed by serine palmitoyltransferase and prevents fumonisin- induced sphinganine accumulation, was used as a tool to further dissect the roles of complex sphingolipid depletion and of sphinganine accumulation on the ERK pathway. As shown in Figure 4.5, the total amount of ERK protein was not affected by any treatments (A). However, phosphorylation of ERK and MEK (B and C, respectively) was reduced in cells cultured with 50 1.1M fumonisin B, for 16 hr. Importantly, addition of 50 1.1M B- fluoroalanine alone did not reduce phosphorylation of ERK or MEK and B-fluoroalanine added together with 50 BM fumonisin B, for 16 hr restored phosphorylated ERK and MEK. Phosphorylation of Raf-1 (D) was not affected by any of the treatments. Sphinganine significantly reduces phosphorylation of ERK and MEK, but not Raf-1. To further investigate the possibility that sphinganine mediates the fumonisin-induced decrease in ERK phosphorylation, LLC-PK, cells were cultured in the absence or presence of a growth inhibitory and apoptotic concentration of D-erythro-sphinganine. Since ceramide is also growth inhibitory and apoptotic, a subtoxic and non-apoptotic dose of fumonisin B, (10 pM) was used to prevent acylation of sphinganine to form ceramide. As shown in Figure 4.6, 10 1.1M fumonisin B, alone had no effect; whereas, 20 |.1.M sphinganine, either alone or in the presence of 10 BM fumonisin B, potently reduced phosphorylation of 96 BF A - - + + HF "‘- a, ”m???- an. nuns-w- Figure 4.5. Fumonisin B, reduces phosphorylated ERK and MEK, but not Raf-l. LLC-PK, cells were cultured in the absence or presence of fumonisin B, (FB,) at 50 pM and/or B—fluoroalanine (BFA) at 50 uM. After 16 hr, cellular protein was isolated and used for Western blot analysis as described under “Materials and Methods”. Above blots are probed against total ERK (A), phosphorylated ERK (B), phosphorylated MEK (C) and phosphorylated Raf-l (D). 5min 30min Con SA FB SA/FB Con SA FB SA/FB A ERK if “'3 M ff? E“? P”: rim-s ems .‘ r3 fih—‘flm B pERK «...-33223; C pMEK -‘*~-mm~m “*mfiwww“ D pRaf -------- Figure 4.6. Sphinganine reduces phosphorylated ERK and MEK, but not Raf-1. LLC-PK, cells were cultured in the presence of vehicle (Con), sphinganine at 20 pM (SA), firmonisin B, at 10 uM (PB), and sphinganine (20 11M) together with fumonisin B, (10 pM)(SA/FB). After 5 or 30 min, cellular protein was isolated and used for Western blot analysis as described under “Materials and Methods”. Above blots are probed against total ERK (A), phosphorylated ERK (B), phosphorylated MEK (C) and phosphorylated Raf-1 (D) 97 ERK (B) and MEK (C) without affecting the total amount of ERK (A). The effects of sphinganine persisted ~30 min, and subsided by ~l hr (data not shown). Raf-1 was not affected by any treatment. F umanisin B, suppresses Akt phosphorylation. To determine the effect of fumonisin B, on the Akt pathway and to dissect the role of sphinganine accumulation, Western blot analysis was conducted using antibodies to total Akt and to Akt phosphorylated at thr3 08 and at ser473. As shown in Figure 4.7, total Akt (C) was not affected by any treatment. However, culture with 50 BM fumonisin B, for 16 hr reduced both pAkt (ser473) (A) and pAkt (thr308) (B). Culture with B-fluoroalanine alone had no effect and addition of B- fluoroalanine together with alone or in the presence of fumonisin B, restored pAkt (thr308) and pAkt (ser473) to that of controls. Sphinganine significantly reduces phosphorylation of A kt. T o d e t e r m i n e i f sphinganine has similar effects as fumonisin B, on Akt phosphorylation, LLC-PK, cells were cultured in the absence or presence of sphinganine. Again, a subtoxic and non-apoptotic concentration of fumonisin B, (10 1.1M) was used to prevent acylation of sphinganine to form ceramide. Cultures of LLC-PK, cells with 10 11M fumonisin B, had no effect; whereas, culture with 20 |J.M sphinganine both alone and in the presence of 10 BM fumonisin B, reduced the phosphorylation of Akt at both thr308 (A) and ser473 (B) at both 5 or 30 min of culture (Figure 4.8). E. DISCUSSION Inhibition of ceramide synthase by fumonisin B, causes accumulation of sphingoid bases and depletion of complex sphingolipids (Norred et al., 1992; Wang et al., 1991; Schroeder 98 FB, - + - + [SPA - - + + A pAkt-ser473 ...... ...... “ms-ml gm. --v----. ,. B pAkt-thr308 ' ...... W .. ...... 4 mm —— ‘i C Akt ----i Figure 4. 7. Fumonisin B, reduces phosphorylated Akt. LLC-PK, cells were cultured in the absence or presence of fumonisin B, (F B,) at 50 11M and/or B—fluoroalanine (BFA) at 50 11M. After 16 hr, cellular protein was isolated and used for Western blot analysis as described under “Materials and Methods”. Above blots are probed against phosphorylated Akt at serine 473 mAkt-ser473) (A) and threonine 308 (pAkt-thr308)(B), and total Akt (C). 5 min 30 rnin Con SA FB SA/FB Con SA FB SA/FB A pAkt‘tIlr308 a" “ “'4 m we» B pAkt-ser473 " "'"' '- "'"""' - ...-... ‘ ~—-' Figure 4.8. Sphinganine reduces phosphorylated Akt. LLC-PK, cells were cultured in the presence of vehicle (Con), sphinganine at 20 uM (SA), fumonisin B, at 10 11M (PB), and sphinganine (20 uM) together with fumonisin B, (10 pM)(SA/FB). After 5 or 30 min, cellular protein was isolated and used for Western blot analysis as described under “Materials and Methods”. Above blots are probed against phosphorylated Akt at serine 473 (pAkt- ser473) (A) and threonine 308 (pAkt-thr308) (B). 99 et al., 1994). Sphingoid bases are growth inhibitory and cytotoxic for many cell types (Merrill et al., 1993b) and complex sphingolipids also have been implicated in the regulation of cell growth and differentiation (Hakomori and I garashi, 1993; Laine and Hakomori, 1973; Bremer and Hakomori, 1982; Saito et al. , 1985). Consistent with previous findings (Kim et al., 2001), fumonisin B, caused morphological changes indicative of apoptosis within 16 hr and inhibited growth of LLC-PK, cells within 24 hr of culture. This coincides with the time when endogenous sphinganine reaches a concentration of ~1 nmol/mg protein which is growth inhibitory and apoptotic (Kim et al., 2001). The potent anti-proliferative effect and concomitant morphological changes characteristic of apoptosis in LLC-PK, cells as early as 1 hr after the addition of exogenous sphinganine support a mechanism of fumonisin toxicity involving sphinganine accumulation. The ability of sphinganine to diffuse rapidly through cell membranes (Hannun et al. , 1991) explains the more rapid action of exogenously added sphinganine. This cytotoxicity of sphinganine in kidney cells is consistent with other reports where sphingoid bases inhibited cell growth and induced apoptosis (Ohta et al. , 1995, 1994; Stevens et al., 1990). The growth inhibitory and cytotoxic effects of sphinganine in LLC- PK, cells explain fumonisin’s interruption of regeneration after mechanical injury and enlargement of swiped area in renal proximal tubule cell culture (Counts et al. , 1995) as well as the apoptotic features shown in the tubular epithelial region of the kidney (Voss et al., 2001; Tolleson et al., 1996a; Bondy et al., 1996; Bucci et al., 1998). In various cell lines including LLC-PK, cells, the ERK pathway plays a critical role in determining cellular growth (di Mari et al., 1999; Dudley et al., 1995; Kinane et al., 1997; Schramek et al., 1997). Phosphorylation and activation of ERK activates transcription factors such as NF -KB and AP-l via a series of signaling events and leads to transcription 100 of genes involved in cell proliferation (Cowley et al., 1994; Hunter, 1995; Marshall, 1995). Previously, fumonisin B, was reported to activate ERK at ~10 min after addition of fumonisin B, to cultures of human bronchial epithelial cells (Pinelli et al., 1999) and Swiss 3T3 cells (Wattenberg et al., 1996). However, in the present study, ERK phosphorylation increased as early as 5-10 min for both control and fumonisin B, cultures and then decreased to the initial level by 30 min, indicating ERK phosphorylation is not rapidly altered in response to fumonisin B ,. The reason for the disparity in results may be because appropriate controls were not included in the previous studies (Pinelli et al., 1999; Wattenberg et al., 1996). Alternatively, the discrepancy in results may be due to distinct metabolic responses to sphingoid bases of different cell types. In contrast to the lack of an early response to fumonisin B,, the present study demonstrates that the ERK pathway was down-regulated within 16 hr of culture with fumonisin B,. Again, this is the point at which endogenous sphinganine has increased to ~l nmol/mg protein (Kim et al, 2001). To our knowledge, this is the first study demonstrating down-regulation of ERK by fumonisin B,. Fumonisin B, also produced similar effects when ERK was stimulated by insulin or inhibited by PD98059, a MEK inhibitor (data not shown). These observations indicate that fumonisin B, inhibits kidney cell proliferation, at least in part, by suppressing the ERK pathway. Further, the time-dependence for inhibition of ERK phosphorylation by fumonisin B, suggests a possible role of sphinganine accumulation. Exogenously added sphinganine also significantly decreased phosphorylation of ERK and MEK, but even more rapidly than fumonisin B,. This is consistent with previous studies which demonstrated that potent inhibition of ERK plays an important role in sphingoid base- mediated death of U937 human monoblastic leukemia cells (Jarvis et al., 1997). A causal 101 role for sphinganine in mediating suppression of the ERK pathway by fumonisin B, is further supported by the finding that addition of B-fluoroalanine (which prevents sphinganine accumulation caused by fumonisin B,) abolished the inhibiting effect of fumonisin B, on phosphorylation of ERK and MEK. The possibility of complex sphingolipid depletion as a cause for firmonisin’s effect on the ERK pathway is unlikely because B-fluoroalanine alone, which depletes complex sphingolipid such as sphingomyelin to a similar degree as fumonisin B, (Merrill et al. , 1993a), did not decrease ERK or MEK phosphorylation. ERK and MEK phosphorylation were inhibited to a similar extent by both sphinganine alone or sphinganine together with a suboptimal concentration of fumonisin B, (which prevents formation of ceramide from exogenous sphinganine), indicating that sphinganine, not ceramide, inhibits phosphorylation of ERK and MEK. Furthermore, neither fumonisin B, nor sphinganine altered phosphorylation of Raf-1 , a serine threonine kinase which mediates signaling from Ras to MEK (Dent et al., 1992; Howe et al., 1992; Kyriakis et al., 1992). These findings demonstrate that down-regulation of ERK phosphorylation by fumonisin B, is mediated by sphinganine accumulation via a mechanism which is independent of Raf-l and which may involve MEK kinase, the other MEK activator. Fumonisin B, and sphinganine may regulate MEK and ERK via MEK kinase by either inhibiting PKC or by affecting heterodimeric G-protein-coupled receptors. PKC regulates the ERK pathway via either MEK kinase or Raf-1 (Hill et al., 1993; Nishida and Gotoh, 1993; Pelech et al., 1993). Furthermore, sphingoid bases are potent inhibitors of PKC (Hannun and Bell, 1986) and fumonisin B, has been shown to inhibit PKC activity in a concentration-dependent manner (Huang et al., 1995). Alternatively, fumonisin B, and sphingoid bases also interact with G-protein-coupled receptors (Ho et al., 1996) and G- 102 protein-coupled receptors regulate ERK activity via MEK kinase, rather than Raf-1 (Cobb and Goldsmith, 1995; Lavoie et al., 1996). Moreover, in LLC-PK, cells Ga,_, has been shown to mediate growth via Raf-independent activation of ERK (Kinane et al., 1997). Fumonisin B, and sphinganine may regulate MEK and ERK via MEK kinase by either inhibiting PKC or by affecting heterotrimeric G protein-coupled receptors. Akt (also known as protein kinase B), a serine/threonine kinase activated by P13 K, plays an important role in the regulation of cell death (Krasilnikov, 2000; Zhou et al., 2000). Akt kinase protects cells from apoptosis induced by a variety of extracellular stresses including loss of cell adhesion, serum withdrawal, and UV-B irradiation (Kennedy et al. , l 997; Khwaja et al., 1997; Kulik et al., 1997). Further, increased activity of Akt contributes to carcinogenesis in various tissues (Graff et al., 2000; Krasilnikov, 2000; Tsatsanis and Spandidos, 2000; Yuan et al., 2000). Recently, Ramljak et al. (2000) showed Akt was activated in a neoplastic liver specimen obtained from a long-term feeding study of fumonisin B, in rats and suggested that fumonisin-induced Akt activation may be a mechanism by which fumonisin B, acts as a carcinogen. In contrast, the present study demonstrates down-regulation of Akt phosphorylation by fumonisin B,. The disparity in results could be due to differences in species (rat vs pig), organ (liver vs kidney), model system (in viva vs. in vitra), and/or study period (chronic vs. acute). Treatment of LLC-PK, cells with PI3K inhibitor LY294002 significantly reduced Akt phosphorylation and the metabolic capacity (data not shown), indicating the importance of Akt in normal LLC-PK, cellular metabolism. F umonisin-induced decreases in Akt phosphorylation may suppress the cell survival signal and shift the balance toward cell death, providing an additional mechanism by which fumonisin B, inhibits cell growth and induces apoptosis. 103 Similar to the ERK pathway, the mechanism by which fumonisin B, down-regulates Akt phosphorylation appears to involve sphinganine accumulation rather than complex sphingolipid depletion. Exogenously added sphinganine significantly reduced phosphorylated Akt on both ser473 and thr308 and B-fluoroalanine prevented the fumonisin- induced decreases in Akt phosphorylation. The possibility that complex sphingolipid depletion mediates the effect of fumonisin B, on Akt phosphorylation is unlikely because B- fluoroalanine alone, which depletes complex sphingolipids but without causing sphinganine accumulation, did not decrease Akt phosphorylation at either ser 473 nor thr308. Down- regulation of Akt and induction of cell death by sphinganine are consistent with those of Chang et al. (2001) who demonstrated that sphingosine, another sphingoid base, induces dephosphorylation of Akt during apoptosis in human hepatoma (Hep3B) cells. Sphingosine also caused cytochrome C release and caspase-3 activation (Chang et al., 2001) which are critical events in the regulation of apoptosis (Kluck et al., 1997). Attenuation of sphingosine-induced cytochrome C release and caspase 3 activation by overexpression of activated Akt kinase (Chang et al. , 2001) implies that suppression of Akt by sphingoid bases could cause cell death. Thus, the cytotoxic effect of sphinganine in the present study is, at least in part, due to the suppression of signals via the PI3K/Akt pathway and this property appears to contribute to the growth-inhibitory and cytotoxic effects of fumonisin B, in LLC- PK, cells, leading to fumonisin-induced kidney cell death. Taken together, the findings of this study suggest that fumonisin-induced growth inhibition and cell death in LLC-PK, porcine renal tubular epithelial cells is mediated by sphinganine which suppresses ERK and Akt signaling. Moreover, down-regulation of the ERK phosphorylation cascade is independent of Raf-1. 104 CHAPTER V SUMMARY AND CONCLUSIONS 105 The present study investigated how fumonisin B,, the most common contaminant of corn, can cause toxicity in the kidney, the most sensitive target organ, using an in vitro culture system with LLC-PK, porcine renal tubular epithelial cells. LLC-PK, cells cultured with fumonisin B, underwent growth arrest and exhibited increased DNA fragmentation, cell shrinkage, and membrane blebbing indicative of apoptosis similar to renal tubular epithelial cells in viva. Flow cytometric analysis revealed that fumonisin B, produced a 7-fold increase in the number of cells in the sub-Go/G, range which represents apoptotic cells. Therefore, LLC-PK, cells appear to be an excellent model to study the molecular mechanism of toxicity caused by fumonisins. Studies which examined the role of disruption of sphingolipid metabolism in fumonisin-induced apoptosis demonstrated that apoptosis was highly dependent on sphinganine accumulation. Since sphinganine accumulation is also evident in kidney and other tissues of animals fed fumonisins, in viva examination of known targets of sphingoid bases and their l-phosphate metabolites may provide insight into the mechanism of action of these mycotoxins. Fumonisin B, also increased calmodulin mRNA and calmodulin protein, apparently by increasing calmodulin gene transcription via sphinganine l-phosphate. Fumonisin B, and sphinganine l-phosphate are the first two examples of chemical inducers of calmodulin. Thus, both fumonisin B, and sphinganine l-phosphate provide novel tools to study transcriptional regulation of calmodulin genes. Since apoptosis induced by fumonisin B, was dependent on calmodulin activity, overexpression of the calmodulin gene either by activating the calmodulin gene promoter or via calmodulin gene transfection may lead to a better understanding of the mechanism of 106 action of fumonisins. In addition, calmodulin antisense cDNA and calmodulin antagonists as well as calmodulin knockout animals may be useful tools to further elucidate the mechanism of fumonisin toxicity. Calmodulin antisense cDNA and calmodulin antagonists also may serve as drugs to treat animals and/or humans exposed to fumonisins. Moreover, calmodulin mRN A and calmodulin protein may provide sensitive biomarkers for exposure to the mycotoxin. Sphinganine l-phosphate appears to mediate fumonisin-induced calmodulin induction and, thereby, contributes to fumonisin-induced apoptosis. This observation is contradictory to dogma which contends that sphinganine l-phosphate is “mitogenic” as opposed to “apoptotic.” The full growth inhibitory and apoptotic response to fumonisin B, appears to be dependent not only on sphinganine l-phosphate and calmodulin, but also on sphinganine and regulation of key signaling pathways. The results underscore that changes in cellular behavior (in this case apoptosis) are the result of a number of mediators acting on a variety of targets rather than a single mediator acting on only one or very few targets. Fumonisin B, decreased phosphorylation of both ERK and Akt proteins via sphinganine accumulation. The more potent effects on ERK and Akt of exogenous sphinganine than fumonisin B, may be due to accumulation of a different profile of mediators. Exogenous sphinganine causes accumulation of sphinganine and sphinganine 1- phosphate as well as ceramide; whereas, fumonisin B, prevents formation of ceramide and causes accumulation of sphinganine and sphinganine l-phosphate. Ceramide has been shown to inhibit ERK and Akt, while sphinganine l-phosphate has been reported to activate ERK and the effect on Akt is not know. The molecular mechanism by which fumonisin B, suppresses ERK and Akt signalng 107 may involve sphinganine inhibition of protein kinase C (PKC). Though the effects of sphinganine and the mycotoxin on PKC were not measured in this study, PKC is an upstream regulator of ERK and sphingoid bases potently inhibit PKC. Future studies should examine the role of PKC in fumonisin-induced suppression of ERK. If inhibition of PKC does play a role, activators of PKC may serve as drugs to counteract the action of fumonisin B, and/or sphinganine There are interconnections between sphinganine l-phosphate and calmodulin which implicate these molecules not only in fumonisin-induced apoptosis, but also in carcinogenesis caused by fumonisins. Sphinganine l-phosphate induces calmodulin expression and mobilizes intracellular calcium and calmodulin is dependent on calcium for some regulatory activities. In addition, the concentration of both sphinganine 1-phosphate and calmodulin appear to be increased by common stimuli including serum and neural growth factor and both have been shown to stimulate DNA synthesis, expedite cell cycle progression through G,/S, and play roles in activation of cyclin-dependent kinases. Moreover, calmodulin is elevated in various forms of tumors. These facts suggest that sphinganine l-phosphate and calmodulin may be key not only in the initial acute response to fumonisin exposure which is apoptosis, but also in carcinogenesis which is the result of chronic exposure to fumonisins. Therefore, future studies should examine the roles of sphinganine l-phosphate and calmodulin in regenerative growth of surviving cells surrounding tissue initially killed by fumonisin exposure, which could contribute to fumonisin-induced carcinogenesis in various tissues including kidney and liver. The findings of this study, taken together with the results of other investigations, provide a plausible molecular mechanism for renal toxicity caused by fumonisins (Figure 108 5.1). Under conditions of exposure to fumonisins via the diet, ceramide synthase is inhibited causing depletion of complex sphingolipids and accumulation of sphinganine and sphinganine l-phosphate. The specific contributions of depletion of complex sphingolipids to fumonisin-induced apoptosis are not clear. Accumulation of sphinganine inhibits P13K and blocks ERK signaling probably by inhibiting protein kinase. Accumulation of sphinganine l-phosphate stimulates calmodulin gene transcription and calcium mobilization which act to increase CaH/CAM and activate the phosphatase calcineurin. Inactivation of ERK and Akt and activation of calcineurin converge to cause dephosphorylation of Bad which triggers apoptosis. Kidney Cell Fumonisin B, /spi?$:lj?;jdsl 5 Ceramide )phinganinA-D Sphinganine 1 Pf Serine + 5 ‘ ‘ V E Palmitoyl CoA . CaM mRNA1 v ea'M [Ca’tlf ‘ 2",; I Ca”ICaM 3‘ Other Enzymes? poptosrs Cell Proliferation I Figure 5.1. Summary of possible molecular mechanism of fumonisin-induced kidney toxicity. (Solid line with arrow, conversion; dotted line with perpendicular line, inhibition; and dotted line with arrow, activation/stimulation) 109 APPENDIX 110 Appendix 1. Primer pairs for amplification of bcl-2. bcl-x and box mRN A (Chapter II) bcl—2 mRNA Forward 5'-ACTTGTGGCCCAGATAGGCACCCAG-3’ Reverse 5'-CGACTTCGCCGAGATGTCCAGCCAG-3' bax mRNA Forward 5'-CAGCTCTGAGCAGATCATGAAGACA-3' Reverse 5’-GCCCATCTTCTTCCAGATGGTGAGC-3' bcl-x mRN A Forward 5'-AATGTCTCAGAGCAACCGGGAGCTG-3' Reverse 5'-TCATITCCTACTGAAGAGTGAGCCCA-3' lll Appendix 2. Differential display RT-PCR. C F R: : 5;” ”I, L'— i-‘t' <-K18 K14 —>‘--»~5°~ K15—>...... ‘ K16—"'""“ :’*~ r;- TAIO K17--" TA6 C, Control; F, 50 uM fumonisin B, Primers: TA6; 5'-T12-A-3' and 5'-GCAATCGATC-3' TA10; 5'-T12-A-3' and 5'-TAGCAAGTGC-3' 112 Sequences of cDNAs K 12 l GAATCNCCNT NGCTTGGGTA CCGAGCTCGG ATCCACTAGT AACGGGCCGC CAGTGTGCTG 61 GAATTCGCCC TTGCAATCGA TGTAACATTA TTATTGCCAC ACACATTTTT ATACTAGATC 121 CTAACACTTC GTCTACATTG CTTTTAAATA ATGTTCGTAT GTCACTGTGT GTTATAAATC 181 TGCATCAATT TCCGTTTTAA TACCAGTATC CTTTTTCCTC TATACGAATC CCTTGAATCT 241 TAGGGTAATG TACCTTACAA AATTTTAGGT TCTACAGATG CCATATTGTC TGAATCTCAC AAGGATTTCA ATATGGCTAT 7! '63 1 ATTGGGCCCT CTAGATGCAT GCTCGAGCGG CCGCCAGTGT GATGGATATC TGCAGAATTC 61 GNCCTTGCAA TCGATGACGG CAAAGAGCAT TGTGATCAAA TCCACATGAG TAAAGTAACA 121 TTTGGAGCCC AGCAGGGGAT AAAAAGGAAG GTATTGTTGC TCGGAGTGCC ACAGTATTGT 181 CCTTTACCAT TTTAGGATGT ATTGAAATGC ATAAGTAATA GTTGTAATTG CATATATGAG 241 AAAGCACCAA GCAAGCCCTA GCTCCTTTGA AAAGGCCGGG ' CTGGGAGGTG AAAATCATAA 301 ATCTGGGCGC CTTGAAACAG CATTCTITAG GGAAGGAGAC TATGTGATCG CTGTGCAGCA 361 GGTTCGAATC CATAATAAGA CAGATCAGAA TCTGAGCCAC TTGAGATGGA TAGGCTCACT 421 TGTGCATACT ACATCGATTG CAAGGGCGAA TTCCAGCACA CTGGCGGCCG TACTAGTGGA 7! Z l CGGGCGAATT GGGCCCTCTA GATGCATGCT CGAGCGGCCG CCAGTGTGAT GGATATCTGC 61 AGAATTCGTC CTTGCAATCG ATGGAAATCT GAATGATGGA TGTCAAAGAG CACTGCAGGC 121 AACGGGGGAC AATTATTTAC CTAGAAAATA TCTACAGTAG ACACGTTCGT CTGTAGTATC 181 TAAAGGTCTT AATTCTGCTG CCGCTCTGCC ACTTACTGCA TGAATCTGTG TGGGTTGAAT 241 TCATTGCTCT GAGGTTTCAT TTGTCCATTT AAAATGCATG ATCTGCATTA TCCACCTCAG 113 7! u- 7: ex 7': \1 301 GAGCTTGTAT TGAAGCTCAA GTGTCAAAAA TGCATCAAAA TGCTAAAGCG TTACCATTAT 361 AAAATGACTG AAATTGTTAC TCATCGATTG CAAGGGCGAA TTCCAGCACA CTGGCGGCCG 1 ACCCTTNGCT TGGTACCCGA GCTCGGATCC ACTAGTAACG GGCCGCCAGT GTGCTGGAA 61 TTCGCCCTT GCAATCGATG TATAAAGACA NTTTTAATAC TCTGCAAATT CAGGGAGTTT 121 TTTCTCCTAT CTCTAGTATC TTTTCACTAT CCAATTACAT CAAAGCCACC CCTGAGCACT 181 GCTCCCTTCA TATTGNCATT GTGGTAACTT ACTTTAAAAC TTCTTTAGAA TAAAGAGGTG 241 ATATGTGTGC AAATTAGCTT TCATCTGGAC CTCAAGGGAG 1 ACNCTTNCTA CTATCCGGCG AATTGNGCCC TCTAGATGCA TGCTCGAGCG GCCGCCAGTG 61 TGATGGATAT CTGCAGAATT CGCCNTTGCA ATCGATGGGC AGGCACGCCA GCCAGACTTC 121 TTAAGAATCC TGGCAAGAGT TAATACAACA GAAGAAACTG AGAACCATAA CTAGCAGCTG 181 TGTCTTTCCT GATTTGTTTG GGTTTTTTAG TAGCCGGCAT ATAATTATGT CTTCATCTTT 241 CAAGAACTGT CATTTTGTCA GTTGCCCAGT TTGATTTATG TGATAAAGNT GAGGATTTGG 301 AAACATACAA TTCAGTTCTA TTITTATCAT GTCAAAGAAT TAAATTAATT AAAACAATTG l CCGGGCGAAT TGGGCCCTCT AGATGCATGC TCGAGCGGCC GCCAGTGTGA TGGATATCTG 61 CAGAATTCGN CCTTGCAATC GATGCAGATC CACAGGCTAA TGCTACCTCT GGTTTAGGAA 121 TAGAGGAAGT AAATTATTCT ACCTATAATC TATTGGAGCA CAGTGCAGAT GTGAAAGATT 181 GTATCCAGAA AACTTCTTCT CCAAATTTAG ACCTTGTACC TTCTCATATT GATTTGGTAG 241 CAGCGGAAAT TGAGTTAGTA GACCGTGACA AGAGAGAATA TATGCTTAAA AAAGCATTAG 301 AAGAGGTGAA ATCTGAGTAT GACTACATCA TCATCGATTG CAAGGGCGAA TTCCAGCACA 114 7\' co 1 TTGGGCCCTC TAGATGCATG CTCGAGCGGC CGCCAGTGTG ATGGATATCT GCAGAATTCG 61 TCCTTTAGCA AGTGCTGCAG CTGACAAAAT CCCCGGGTTG TTAGGTGTCT TTCAGAAGCT 121 GATTGCATCC AAAGCCAATG ACCACCAAGG TI I I IATCTT CTAAACAGTA TAATAGAGCA 181 CATGCCTCCT GAGTCAGTTG ACCAGTCAGG AAGCAAATCT TCATTCTGCT ATTCCAAAGA 241 CTTCAGAATT CCAAAACAAC CAAGTTTATC AAGAGI 1 ICT TAGTCTTTAT TAATTTGNAT 301 TGCATAAAAT ACGGGGCACT AGCACTGCAA GAAATAI I IG ACGGTATACA ACCAAAAATG 361 GTTGGAATGG GI I IGGAAAA AATCATTATT CCTGGAATTC AGAAAGTATC TGGGAATGTA 115 REFERENCES 116 REFERENCES Ahn, E-H, and Schroeder, H. (2002) Bioactive sphingolipids are constituents of soy and dairy products. J. Food Sci. (in press) Ames, B. N., and Gold, L. S. (1990) Too many rodent carcinogens: mitogenesis increases mutagenesis. Science 249,970-1 Attardi, L. D., Lowe, S. W., Brugarolas, J ., and Jacks, T. (1996) Transcriptional activation by p53, but not induction of the p21 gene, is essential for oncogene-mediated apoptosis. EMBOJ 15,3693-701 Bai, G., and Weiss, B. (1991) The increase of calmodulin in PC 12 cells induced by NGF is caused by differential expression of multiple mRNAs for calmodulin. J Cell Physiol 149,414-21. Bargou, R. C., Daniel, P. T., Mapara, M. Y., Bommert, K., Wagener, C., Kallinich, B., Royer, H. D., and Dorken, B. (1995b) Expression of the bcl-2 gene family in normal and malignant breast tissue: low bax-alpha expression in tumor cells correlates with resistance towards apoptosis. Int J Cancer 60,854-9 Bargou, R. C., Bommert, K., Weinmann, P., Daniel, P. T., Wagener, C., Mapara, M. Y., and Dorken, B. (19953) Induction of Bax-alpha precedes apoptosis in a human B lymphoma cell line: potential role of the bcl-2 gene family in surface IgM-mediated apoptosis. Eur J Immunol 25,770-5 Basu, S., Bayoumy, S., Zhang, Y., Lozano, J., and Kolesnick, R. (1998) BAD enables ceramide to signal apoptosis via Ras and Raf-1. J Biol Chem 273,30419-26. Bell, R.M., Hannun, Y.A., and Merrill, A.H., Jr. (Eds) (1993) Advances in Lipid Research.“ Sphingolipids and Their Metabolites, Vols. 25 and 26. Academic Press, Orlando, FL. Bellomo, G., Perotti, M., Taddei, F., Mirabelli, F., Finardi, G., Nicotera, P., and Orrenius, S. (1992) Tumor necrosis factor alpha induces apoptosis in mammary adenocarcinoma cells by an increase in intranuclear free Ca2+ concentration and DNA fragmentation. Cancer Res 52,1342-6. Bielawska, A. E., Shapiro, J. P., Jiang, L., Melkonyan, H. S., Piot, C., Wolfe, C. L., Tomei, L. D., Hannun, Y. A., and Umansky, S. R. (1997) Ceramide is involved in triggering of cardiomyocyte apoptosis induced by ischemia and reperfusion. Am J Pathol 151,1257-63 Bito, H., Deisseroth, K., and Tsien, R. W. (1997) Cay-dependent regulation in neuronal gene expression. Curr Opin Neurobiol 7,419-29. 117 Blank, M. L., Cress, E. A., Smith, Z. L., and Snyder, F. (1992) Meats and fish consumed in the American diet contain substantial amounts of ether-linked phospholipids. J Nutr 122,1656-61. Bondy, G., Barker, M., Mueller, R., Fernie, S., Miller, J. D., Armstrong, C., Hierlihy, S. L., Rowsell, P., and Suzuki, C. (1996) F umonisin B, toxicity in male Sprague-Dawley rats. Adv Exp Med Biol 392,251-64. Bomfeldt, K. E., Graves, L. M., Raines, E. W., Igarashi, Y., Wayman, G., Yamamura, S., Yatomi, Y., Sidhu, J. S., Krebs, E. G., Hakomori, S., and et al. (1995) Sphingosine-l- phosphate inhibits PDGF-induced chemotaxis of human arterial smooth muscle cells: spatial and temporal modulation of PDGF chemotactic signal transduction. J C ell Biol 130,193-206 Bouvet, M., Bold, R. J., Lee, J ., Evans, D. B., Abbruzzese, J. L., Chiao, P. J., McConkey, D. J., Chandra, J ., Chada, S., Fang, B., and Roth, J. A. (1998) Adenovirus-mediated wild-type p53 tumor suppressor gene therapy induces apoptosis and suppresses growth of human pancreatic cancer. Ann Surg Oncol 5,681 -8 Bremer, E. G., and Hakomori, S. (1982) GM3 ganglioside induces hamster fibroblast growth inhibition in chemically-defined medium: ganglioside may regulate growth factor receptor function. Biochem Biophys Res Commun 106,711-8. Bruckheimer, E. M., Cho, S. H., Sarkiss, M., Hemnann, J ., and McDonnell, T. J. (1998) The Bcl-2 gene family and apoptosis. Adv Biochem Eng Biotechnal 62,75-105 Bucci, T. J., Howard, P. C., Tolleson, W. H., Laborde, J. B., and Hansen, D. K. (1998) Renal effects of fumonisin mycotoxins in animals. Toxicol Pathol 26,160-4 Cawood, M. E., Gelderblom, W. C., Alberts, J. F., and Snyman, S. D. (1994) Interaction of l4C-labelled fumonisin B mycotoxins with primary rat hepatocyte cultures. Food Chem Toxicol 32,627-32. Chafouleas, J. G., Lagace, L., Bolton, W. E., Boyd, A. E., 3rd, and Means, A. R. (1984) Changes in calmodulin and its mRNA accompany reentry of quiescent (G0) cells into the cell cycle. Cell 36,73-81. Chafouleas, J. G., Bolton, W. E., Hidaka, H., Boyd, A. E., 3rd, and Means, A. R. (1982) Calmodulin and the cell cycle: involvement in regulation of cell-cycle progression. Cell 28,41-50. Chafouleas, J. G., Pardue, R. L., Brinkley, B. R., Dedman, J. R., and Means, A. R. (1981) Regulation of intracellular levels of calmodulin and tubulin in normal and transformed cells. Prac Natl Acad Sci U S A 78,996-1000. Chang, H. C., Tsai, L. H., Chuang, L. Y., and Hung, W. C. (2001) Role of AKT kinase in sphingosine-induced apoptosis in human hepatoma cells. J Cell Physiol 188,188-93. 118 Chao, R., Khan, W., and Hannun, Y. A. (1992) Retinoblastoma protein dephosphorylation induced by D-erythro-sphingosine. J Biol Chem 267,23459-62 Charles, A. G., Han, T. Y., Liu, Y. Y., Hansen, N., Giuliano, A. E., and Cabot, M. C. (2001) Taxol-induced ceramide generation and apoptosis in human breast cancer cells. Cancer C hemather Pharmacol 47 ,444-50. Chen, J ., Nikolova-Karakashian, M., Menill, A. H., Jr., and Morgan, E. T. (1995) Regulation of cytochrome P450 2C11 (CYP2C11) gene expression by interleukin-1, sphingomyelin hydrolysis, and ceramides in rat hepatocytes. J Biol Chem 270,25233-8. Chomczynski, P., and Sacchi, N. (1987) Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroforrn extraction. Anal Biochem 162,156-9 Ciacci Zanella, J. R., and Jones, C. (1999) F umonisin B,, a mycotoxin contaminant of cereal grains, and inducer of apoptosis via the tumour necrosis factor pathway and caspase activation. F aad Chem T axicol 37,703-12 Cifone, M. G., De Maria, R., Roncaioli, P., Rippo, M. R., Azuma, M., Lanier, L. L., Santoni, A., and Testi, R. (1994) Apoptotic signaling through CD95 (Fas/Apo-l) activates an acidic sphingomyelinase. J Exp Med 180,1547-52. Cobb, M. H., and Goldsmith, E. J. (1995) How MAP kinases are regulated. J Biol Chem 270, 14843- 6. Cohen, S. M., and Ellwein, L. B. (1990) Cell proliferation in carcinogenesis [see comments]. Science 249,1007-1 l Colvin, B. M., and Harrison, L. R. (1992) Fumonisin-induced pulmonary edema and hydrothorax in swine. Mycapathologia 117,79-82 Conway, A. M., Pyne, N. J ., and Pyne, S. (1997) Sphingosine l-phosphate activation of MAP kinase--involvement of P1 3-kinase and protein kinase C. Biochem Soc Trans 25,S585 Counts, R. S., Nowak, G., Wyatt, R. D., and Schnellmann, R. G. (1995) Nephrotoxicant inhibition of renal proximal tubule cell regeneration. Am J Physiol 269,F274-8l. Cowley, S., Paterson, H., Kemp, P., and Marshall, C. J. (1994) Activation of MAP kinase kinase is necessary and sufficient for PC 12 differentiation and for transformation of NIH 3T3 cells. Cell 77,841-52. Cuvillier, O., Rosenthal, D. S., Smulson, M. E., and Spiegel, S. (1998) Sphingosine 1- phosphate inhibits activation of caspases that cleave poly(ADP-ribose) polymerase and larnins during Fas- and ceramide-mediated apoptosis in Jurkat T lymphocytes. J Biol Chem 273,2910-6 119 Cuvillier, O., Pirianov, G., Kleuser, B., Vanek, P. G., C080, 0. A., Gutkind, S., and Spiegel, S. (1996) Suppression of ceramide-mediated programmed cell death by sphingosine-1- phosphate. Nature 381 ,800-3. Davaille, J., Gallois, C., Habib, A., Li, L., Mallat, A., Tao, J., Levade, T., and Lotersztajn, S. (2000) Antiproliferative properties of sphingosine l-phosphate in human hepatic myofibroblasts. J Biol Chem 275, 34628-34633 Davis, B. H., Chen, A., and Beno, D. W. (1996) Raf and mitogen-activated protein kinase regulate stellate cell collagen gene expression. J Biol Chem 271,11039-42. Dawson, G., Goswami, R., Kilkus, J ., Wiesner, D., and Dawson, S. (1998) The formation of ceramide from sphingomyelin is associated with cellular apoptosis. Acta Biochim Pol 45,287-97 Dbaibo, G. S., Wolff, R. A., Obeid, L. M., and Hannun, Y. A. (1995) Activation of a retinoblastoma-protein-dependent pathway by sphingosine. Biochem J 310,453-9 De Jonghe, S., Van Overrneire, 1., Poulton, S., Hendrix, C ., Busson, R., Van Calenbergh, S., De Keukeleire, D., Spiegel, S., and Herdewijn, P. (1999) Structure-activity relationship of short-chain sphingoid bases as inhibitors of sphingosine kinase. Bioorg Med Chem Lett 9,3175-80. Dent, P., Haser, W., Haystead, T. A., Vincent, L. A., Roberts, T. M., and Sturgill, T. W. (1992) Activation of mitogen-activated protein kinase kinase by v-Raf in NIH 3T3 cells and in vitro. Science 257,1404-7. Desai, N. N., Zhang, H., Olivera, A., Mattie, M. E., and Spiegel, S. (1992) Sphingosine-1- phosphate, a metabolite of sphingosine, increases phosphatidic acid levels by phospholipase D activation. J Biol Chem 267,23122-8 Desbarats, L., Schneider, A., Muller, D., Burgin, A., and Eilers, M. (1996) Myc: a single gene controls both proliferation and apoptosis in mammalian cells. Experientia 52,1123-9 di Mari, J. E, Davis, R., and Safirstein, R. L. (1999) MAPK activation determines renal epithelial cell survival during oxidative injury. Am J Physiol 277,F195-203 Dobrowsky, R. T., and Hannun, Y. A. (1992) Ceramide stimulates a cytosolic protein phosphatase. J Biol Chem 267,5048-51. Donzelli, M., Bemardi, R., Negri, C., Prosperi, E., Padovan, L., Lavialle, C., Brison, O., and Scovassi, A. I. (1999) Apoptosis-prone phenotype of human colon carcinoma cells with a high level amplification of the c-myc gene. Oncogene 18,439-48 Dowd, D. R., MacDonald, P. N., Komm, B. S., Haussler, M. R., and Miesfeld, R. (1991) Evidence for early induction of calmodulin gene expression in lymphocytes undergoing 120 glucocorticoid-mediated apoptosis. J Biol Chem 266,18423-6. Dudley, D. T., Pang, L., Decker, S. J., Bridges, A. J ., and Saltiel, A. R. (1995) A synthetic inhibitor of the mitogen-activated protein kinase cascade. Proc Natl Acad Sci USA 92,7686- 9 Ealovega, M. W., McGinnis, P. K., Sumantran, V. N., Clarke, M. F., and Wicha, M. S. (1996) bcl-x, gene therapy induces apoptosis of human mammary tumors in nude mice. Cancer Res 56,1965-9 Edrington, T. S., Kamps-Holtzapple, C. A., Harvey, R. B., Kubena, L. F., Elissalde, M. H., and Rottinghaus, G. E. (1995) Acute hepatic and renal toxicity in lambs dosed with fumonisin-containing culture material. J Anim Sci 73,508-15. Edsall, L. C., Pirianov, G. G., and Spiegel, S. (1997) Involvement of sphingosine 1- phosphate in nerve growth factor-mediated neuronal survival and differentiation. J Neurosci 17,6952-60 el Touny, S., Khan, W., and Hannun, Y. (1990) Regulation of platelet protein kinase C by oleic acid. Kinetic analysis of allosteric regulation and effects on autophosphorylation, phorbol ester binding, and susceptibility to inhibition. J Biol Chem 265,16437-43 Epstein, P. N., Christenson, M. A., and Means, A. R. (1989) Chicken calmodulin promoter activity in proliferating and differentiated cells. Mol Endocrinol 3,193-202. Evain, D., Klee, C., and Anderson, W. B. (1979) Chinese hamster ovary cell population density affects intracellular concentrations of calcium-dependent regulator and ability of regulator to inhibit adenylate cyclase activity. Prac Natl Acad Sci U S A 76,3962-6. F raga, C. G., Shigenaga, M. K., Park, J. W., Degan, P., and Ames, B. N. (1990) Oxidative damage to DNA during aging: 8-hydroxy-2'-deoxyguanosine in rat organ DNA and urine. Prac Natl Acad Sci U S A 87,4533-7. Futerman, A. H., and Pagano, R. E. (1991) Determination of the intracellular sites and topology of glucosylceramide synthesis in rat liver. Biochem J 280,295-302. Galderisi, U., Di Bernardo, G., Cipollaro, M., Peluso, G., Cascino, A., Cotrufo, R., and Melone, M. A. (1999) Differentiation and apoptosis of neuroblastoma cells: role of N-myc gene product. J Cell Biochem 73,97-105 Garcia Bermejo, L., Vilaboa, N. E., Perez, C., Galan, A., De Blas, E., and Aller, P. (1997) Modulation of heat-shock protein 70 (HSP70) gene expression by sodium butyrate in U-937 promonocytic cells: relationships with differentiation and apoptosis. Exp Cell Res 236,268- 74 Gelderblom, W. C., Snyman, S. D., Abel, S., Lebepe-Mazur, S., Smuts, C. M., Van der 121 Westhuizen, L., Marasas, W. F., Victor, T. C., Knasmuller, S., and Huber, W. (1996a) Hepatotoxicity and carcinogenicity of the fumonisins in rats. A review regarding mechanistic implications for establishing risk in humans. Adv Exp Med Biol 392,279-96. Gelderblom, W. C., Cawood, M. E., Snyman, S. D., Vleggaar, R., and Marasas, W. F. (1993) Structure-activity relationships of fumonisins in short-terrn carcinogenesis and cytotoxicity assays. F oad Chem Toxicol 31,407-14 Gelderblom, W. C., Kriek, N. P., Marasas, W. F., and Thiel, P. G. (1991) Toxicity and carcinogenicity of the F usarium moniliforme metabolite, fumonisin B,, in rats. Carcinogenesis 12,1247-5 1 Gelderblom, W. C., Snyman, S. D., Lebepe-Mazur, S., van der Westhuizen, L., Kriek, N. P., and Marasas, W. F. (1996b) The cancer-promoting potential of fumonisin B, in rat liver using diethylnitrosamine as a cancer initiator. Cancer Lett 109,101-8. Gelderblom, W. C., Jaskiewicz, K., Marasas, W. F., Thiel, P. G., Horak, R. M., Vleggaar, R., and Kriek, N. P. (1988) Fumonisins--novel mycotoxins with cancer-promoting activity produced by Fusarium moniliforme. Appl Environ Microbiol 54,1806-11. Ghosh, T. K., Bian, J., and Gill, D. L. (1994) Sphingosine l-phosphate generated in the endoplasmic reticulum membrane activates release of stored calcium. J Biol Chem 269,22628-35 Ghosh, T. K., Bian, J., and Gill, D. L. (1990) Intracellular calcium release mediated by sphingosine derivatives generated in cells. Science 248,1653-6 Goetzl, E. J., and An, S. (1998) Diversity of cellular receptors and functions for the lysophospholipid growth factors lysophosphatidic acid and sphingosine l-phosphate. FASEB J 12,1589-98 Graff, J. R., Konicek, B. W., McNulty, A. M., Wang, Z., Houck, K., Allen, S., Paul, J. D., Hbaiu, A., Goode, R. G., Sandusky, G. E., Vessella, R. L., and Neubauer, B. L. (2000) Increased AKT activity contributes to prostate cancer progression by dramatically accelerating prostate tumor growth and diminishing p27Kip1 expression. J Biol Chem 275,24500-5. Grassilli, E., Carcereri de Prati, A., Monti, D., Troiano, L., Menegazzi, M., Barbieri, D., Franceschi, C ., and Suzuki, H. (1992) Studies of the relationship between cell proliferation and cell death. 11. Early gene expression during concanavalin A-induced proliferation or dexamethasone-induced apoptosis of rat thymocytes. Biochem Biophys Res Commun 188,1261-6 Groves, F. D., Zhang, L., Chang, Y. S., Ross, P. F., Casper, H., Norred, W. P., You, W. C., and Fraumeni, J. F. (1999) F usarium mycotoxins in corn and corn products in a high-risk area for gastric cancer in Shandong Province, China. J AOAC Int 82,657-62. 122 Gumprecht, L. A., Marcucci, A., Weigel, R. M., Vesonder, R. F ., Riley, R. T., Showker, J. L., Beasley, V. R., and Haschek, W. M. (1995) Effects of intravenous fumonisin B, in rabbits: nephrotoxicity and sphingolipid alterations. Nat Toxins 3,3 95-403 Gutema, T., Munimbazi, C., and Bullerrnan, L. B. (2000) Occurrence of fumonisins and moniliforrnin in corn and com-based food products of US. origin. J Food Prat 63,1 732-7. Haimovitz-Friedman, A., Kan, C. C., Ehleiter, D., Persaud, R. S., McLoughlin, M., Fuks, Z., and Kolesnick, R. N. (1994) Ionizing radiation acts on cellular membranes to generate ceramide and initiate apoptosis. J Exp Med 180,525-35. Hakomori, S., and Igarashi, Y. (1993) Gangliosides and glycosphingolipids as modulators of cell growth, adhesion, and transmembrane signaling. Adv Lipid Res 25,147-62. Hakomori, S. (1981) Glycosphingolipids in cellular interaction, differentiation, and oncogenesis. Annu Rev Biochem 50,733-64. Hanada, K., Nishijima, M., Kiso, M., Hasegawa, A., Fujita, S., Ogawa, T., and Akamatsu, Y. (1992) Sphingolipids are essential for the growth of Chinese hamster ovary cells. Restoration of the growth of a mutant defective in sphingoid base biosynthesis by exogenous sphingolipids. J Biol Chem 267,23527-33 Hannun, Y. A. (1994) The sphingomyelin cycle and the second messenger function of ceramide. J Biol Chem 269,3125-8 Hannun, Y. A., and Bell, R. M. (1989) Regulation of protein kinase C by sphingosine and lysosphingolipids. Clin Chim Acta 185,333-45 Hannun, Y. A., and Bell, R. M. (1986) Phorbol ester binding and activation of protein kinase C on triton X-100 mixed micelles containing phosphatidylserine. J Biol Chem 261,9341-7. Hannun, Y. A., and Bell, R. M. (1987) Mechanism of activation of protein kinase C: role of diacylglycerol and calcium second messengers. Sac Gen Physiol Ser 42,229-40 Hannun, Y. A., and Obeid, L. M. (1995) Ceramide: an intracellular signal for apOptosis. Trends Biochem Sci 20,73-7 Hannun, Y. A., Loomis, C. R., Merrill, A. H., Jr., and Bell, R. M. (1986) Sphingosine inhibition of protein kinase C activity and of phorbol dibutyrate binding in vitro and in human platelets. J Biol Chem 261,12604-9 Hannun, Y. A., Merrill, A. H., Jr., and Bell, R. M. (1991) Use of sphingosine as inhibitor of protein kinase C. Methods Enzymal 201,316-28. Harel, R., and Futerman, A. H. (1993) Inhibition of sphingolipid synthesis affects axonal 123 outgrowth in cultured hippocampal neurons. J Biol Chem 268,14476-81 Harrison, L. R., Colvin, B. M., Greene, J. T., Newman, L. E., and Cole, J. R. (1990) Pulmonary edema and hydrothorax in swine produced by fumonisin B,, a toxic metabolite of Fusarium moniliforme. J Vet Diagn Invest 2,217-21. Haschek, W. M., Motelin, G., Ness, D. K., Harlin, K. S., Hall, W. F ., Vesonder, R. F ., Peterson, R. E., and Beasley, V. R. (1992) Characterization of fumonisin toxicity in orally and intravenously dosed swine. Mycopathologia 117,83-96 Heist, E. K., and Schulman, H. (1998) The role of Ca 2+/calmodulin-dependent protein kinases within the nucleus. Cell Calcium 23,103-14. Hill, C. S., Marais, R., John, S., Wynne, J ., Dalton, S., and Treisman, R. (1993) Functional analysis of a growth factor-responsive transcription factor complex. Cell 73,395-406. Himpens, B., De Smedt, H., and Casteels, R. (1994) Relationship between [Ca2+] changes in nucleus and cytosol. Cell Calcium 16,239-46. Hla, T. (2001) Sphingosine l-phosphate receptors. Prastaglandins Other Lipid Mediat 64,135-42. Ho, A. K., Peng, R., Ho, A. A., Duffield, R., and Dombrink-Kurtzrnan, M. A. (1996) Interactions of fumonisins and sphingoid bases with GTP-binding proteins. Biochem Arch 12, 249-260. Howard, P. C., Eppley, R. M., Stack, M. E., Warbritton, A., Voss, K. A., Lorentzen, R. J., Kovach, R. M., and Bucci, T. J. (2001) Fumonisin B, carcinogenicity in a two-year feeding study using F344 rats and b6c3f1 mice. Environ Health Perspect 109,277-82. Howe, L. R., Leevers, S. J ., Gomez, N., Nakielny, S., Cohen, R, and Marshall, C. J. (1992) Activation of the MAP kinase pathway by the protein kinase raf. Cell 71,335-42. Huang, C., Dickman, M., Henderson, G., and Jones, C. (1995) Repression of protein kinase C and stimulation of cyclic AMP response elements by fumonisin, a fungal encoded toxin which is a carcinogen. Cancer Res 55,1655-9. Hunter, T. (1995) Protein kinases and phosphatases: the yin and yang of protein phosphorylation and signaling. Cell 80,225-36. Igarashi, Y., Kitamura, K., Toyokuni, T., Dean, 3., F enderson, B., Ogawass, T., and Hakomori, S. (1990) A specific enhancing effect of N,N-dimethylsphingosine on epidermal growth factor receptor autophosphorylation. Demonstration of its endogenous occurrence (and the virtual absence of unsubstituted sphingosine) in human epiderrnoid carcinoma A43 1 cells. J Biol Chem 265,53 85-9. 124 Ishizuka, S., Yano, T., Hagiwara, K., Sone, M., Nihei, H., Ozasa, H., and Horikawa, S. (1999) Extracellular signal-regulated kinase mediates renal regeneration in rats with myoglobinuric acute renal injury. Biochem Biophys Res Commun 254,88-92 Jarvis, W. D., Fomari, F. A., Jr., Auer, K. L., Freemerrnan, A. J., Szabo, E., Birrer, M. J., _ Johnson, C. R., Barbour, S. 13., Dent, P., and Grant, S. (1997) Coordinate regulation of stress- and mitogen-activated protein kinases in the apoptotic actions ofceramide and sphingosine. Mol Pharmacol 52,93 5-47. Jarvis, W. D., Fomari, F. A., Traylor, R. S., Martin, H. A., Kramer, L. B., Erukulla, R. K., Bittman, R., and Grant, S. (1996) Induction of apoptosis and potentiation of ceramide- mediated cytotoxicity by sphingoid bases in human myeloid leukemia cells. J Biol Chem 271,8275-84 Karlsson, K. A. (1970) On the chemistry and occurrence of sphingolipid long-chain bases. Chem Phys Lipids 5,6-43. Kawai, T., Nomura, F., Hoshino, K., Copeland, N. G., Gilbert, D. J., Jenkins, N. A., and Akira, S. (1999) Death-associated protein kinase 2 is a new calcium/calmodulin-dependent protein kinase that signals apoptosis through its catalytic activity. Oncogene 18,3471-80. Kellerman, T. S., Marasas, W. F., Thiel, P. G., Gelderblom, W. C., Cawood, M., and Coetzer, J. A. (1990) Leukoencephalomalacia in two horses induced by oral dosing of fumonisin B1. Onderstepaart J Vet Res 57,269-75. Kennedy, S. G., Wagner, A. J ., Conzen, S. D., Jordan, J ., Bellacosa, A., Tsichlis, P. N., and Hay, N. (1997) The PI 3-kinase/Akt signaling pathway delivers an anti-apoptotic signal. Genes Dev 11,701-13. Khwaja, A., Rodriguez-Viciana, P., Wennstrom, S., Wame, P. H., and Downward, J. (1997) Matrix adhesion and Ras transformation both activate a phosphoinositide 3-OH kinase and protein kinase B/Akt cellular survival pathway. EMBO J 16,2783-93. Kihara Negishi, F ., Yamada, T., Kubota, Y., Kondoh, N., Yamamoto, H., Abe, M., Shirai, T., Hashimoto, Y., and Oikawa, T. (1998) Down-regulation of c-myc and bcl-2 gene expression in PU. 1 -induced apoptosis in murine erythroleukemia cells. lntJ Cancer 76,523- 30 Kim, M. Y., Linardic, C., Obeid, L., and Hannun, Y. (1991) Identification of sphingomyelin turnover as an effector mechanism for the action of tumor necrosis factor alpha and gamma- interferon. Specific role in cell differentiation. J Biol Chem 266,484-9 Kim, M. S., Lee, D. Y., Wang, T., and Schroeder, J. J. (2001) Fumonisin B, induces apoptosis in LLC-PK, renal epithelial cells via a sphinganine- and calmodulin-dependent pathway. Toxicol Appl Pharmacol 176,118-26. 125 Kim, M. S., and Schroeder, J. J. (2002) Sphinganine l-phosphate mediates calmodulin expression induced by fumonisin B, in LLC-PK, porcine renal tubular epithelial cells. (In preparation) Kinane, T. B., Kang, 1., Chu, A., Chin, S. H., and Ercolani, L. (1997) G alpha(i-Z) mediates renal LLC-PK, growth by a Raf-independent activation of p42/p44 MAP kinase. Am J Physiol 272,F273-82. Kluck, R. M., Bossy-Wetzel, E., Green, D. R., and Newrneyer, D. D. (1997) The release of cytochrome c from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science 275,1132-6. Kolesnick, R., and Golde, D. W. (1994) The sphingomyelin pathway in tumor necrosis factor and interleukin-1 signaling. Cell 77,325-8. Kondo, S., Yin, D., Aoki, T., Takahashi, J. A., Morimura, T., and Takeuchi, J. (1994) bcl-2 gene prevents apoptosis of basic fibroblast growth factor-deprived murine aortic endothelial cells. Exp Cell Res 213,428-32 Krasilnikov, M. A. (2000) Phosphatidylinositol-3 kinase dependent pathways: the role in control of cell growth, survival, and malignant transformation. Biochemistry (Mosc) 65,59- 67. ' Kulik, G., Klippel, A., and Weber, M. J. (1997) Antiapoptotic signalling by the insulin-like growth factor I receptor, phosphatidylinositol 3-kinase, and Akt. Mol Cell Biol 17,1 595-606. Kwon, O. S., Schmued, L. C., and Slikker, W. (1997) Fumonisin B, in developing rats alters brain sphinganine levels and myelination. Neurotaxicology 18,571-79. Kyriakis, J. M., App, H., Zhang, X. F ., Banerjee, P., Brautigan, D. L., Rapp, U. R., and Avruch, J. (1992) Raf-1 activates MAP kinase-kinase. Nature 358,417-21. Laine, R. A., and Hakomori, S. (1973) Incorporation of exogenous glycosphingolipids in plasma membranes of cultured hamster cells and concurrent change of growth behavior. Biochem Biophys Res Commun 54,1039-45. Lange-Carter, C. A., Pleiman, C. M., Gardner, A. M., Blumer, K. J., and Johnson, G. L. (1993) A divergence in the MAP kinase regulatory network defined by MEK kinase and Raf. Science 260,315-9. Lavoie, J. N., Rivard, N., L'Allemain, G., and Pouyssegur, J. (1996) A temporal and biochemical link between growth factor-activated MAP kinases, cyclin D1 induction and cell cycle entry. Prog Cell Cycle Res 2,49-58. Lim, C. W., Parker, H. M., Vesonder, R. F., and Haschek, W. M. (1996) Intravenous 126 fumonisin B, induces cell proliferation and apoptosis in the rat. Nat Toxins 4,34-41. Liu, G. X., Sheng, H. F., and Wu, S. (1996) A study on the levels of calmodulin and DNA in human lung cancer cells. Br J Cancer 73,899-901. Machwate, M., Rodan, S. B., Rodan, G. A., and Harada, S. I. (1998) Sphingosine kinase mediates cyclic AMP suppression of apoptosis in rat periosteal cells. Mol Pharmacol 54,70-7 Marasas, W. F. O., Kellerman, T. S., Gelderblom, W. C. A., Coetzer, J. A. W., Thiel, P. G., and Van Der Lugt, J. J. (1988a) Leukoencephalomalacia in a horse induced by fumonisin B, isolated from F usarium moniliforme. Onderstepaart J Vet Res 55, 197-203. Marasas, W. F. (1996) F umonisins: history, world-wide occurrence and impact. Adv Exp Med Biol 392,1 -1 7. Marasas, W. F., Jaskiewicz, K., Venter, F. S., and Van Schalkwyk, D. J. (1988b) F usarium moniliforme contamination of maize in oesophageal cancer areas in Transkei. S A )9 Med J 74,110-4 Marasas, W. F., Kriek, N. P., F incham, J. E., and van Rensburg, S. J. (1984) Primary liver cancer and oesophageal basal cell hyperplasia in rats caused by F usarium moniliforme. Int J Cancer 34,3 83-7 Marshall, C. J. (1995) Specificity of receptor tyrosine kinase signaling: transient versus sustained extracellular signal-regulated kinase activation. Cell 80,179-85. Martinova, E. A., and Merrill, A. H., Jr. (1995) Fumonisin B, alters sphingolipid metabolism and immune function in BALB/c mice: immunological responses to fumonisin B,. Mycopathologia 130,1 63-70. Mattie, M., Brooker, G., and Spiegel, S. (1994) Sphingosine- 1 -phosphate, a putative second messenger, mobilizes calcium from internal stores via an inositol trisphosphate-independent pathway. J Biol Chem 269,3181-8 McConkey, D. J., Nicotera, P., Hartzell, P., Bellomo, G., Wyllie, A. H., and Orrenius, S. (1989) Glucocorticoids activate a suicide process in thymocytes through an elevation of cytosolic Ca 2‘ concentration. Arch Biochem Biophys 269,365-70. Meivar Levy, 1., and Futerman, A. H. (1999) Up-regulation of neutral glycosphingolipid synthesis upon long term inhibition of ceramide synthesis by filmonisin B,. J Biol Chem 274,4607-12 Meivar Levy, 1., Sabanay, H., Bershadsky, A. D., and Futerman, A. H. (1997) The role of sphingolipids in the maintenance of fibroblast morphology. The inhibition of protrusional activity, cell spreading, and cytokinesis induced by fumonisin B, can be reversed by 127 ganglioside GM3. J Biol Chem 272,1558-64 Merrill, A. H., Jr., Wang, E., Mullins, R. E., Jamison, W. C., Nimkar, S., and Liotta, D. C. (1988) Quantitation of free sphingosine in liver by hi gh-performance liquid chromatography. Anal Biochem 171,373-81 Merrill, A. H., Jr., Schmeiz, E. M., Dillehay, D. L., Spiegel, S., Shayaman, J. A., Schroeder, J. J., Riley, R. T. Voss, K. A., and Wang, E. (1997a) Sphingolipids--the enigmatic lipid class: biochemistry, physiology, and pathophysiology. T axtcol appl pharmacal. Orlando, Fla” Academic Press Inc. Jan 208- 225 Menill, A. H., Jr., Schmelz, E. M., Wang, E., Dillehay, D. L., Rice, L. G., Meredith, F., and Riley, R. T. (1997b) Importance of sphingolipids and inhibitors of sphingolipid metabolism as components of animal diets. J Nutr 127,830s Merrill, A. H., Jr., Nimkar, S., Menaldino, D., Hannun, Y. A., Loomis, C., Bell, R. M., Tyagi, S. R., Lambeth, J. D., Stevens, V. L., Hunter, R., and et al. ( 1989) Structural requirements for long-chain (sphingoid) base inhibition of protein kinase C in vitro and for the cellular effects of these compounds. Biochemistry 28,3138-45 Merrill, A. H., Jr., van Echten, G., Wang, E., and Sandhoff, K. (1993a) Fumonisin B, inhibits sphingosine (sphinganine) N-acyltransferase and de nova sphingolipid biosynthesis in cultured neurons in situ. J Biol Chem 268,27299-306. Merrill, A. H., Jr., and Wang, E. (1986) Biosynthesis of long-chain (sphingoid) bases from serine by LM cells. Evidence for introduction of the 4-trans-double bond after de novo biosynthesis of N-acylsphinganine(s). J Biol Chem 261,3764-9. Menill, A. H., Jr., Hannun, Y. A., and Bell, R. M. (1993b) Introduction: sphingolipids and their metabolites in cell regulation. Adv Lipid Res 25,1 -24. Merrill, A. H., Jr., Sereni, A. M., Stevens, V. L., Hannun, Y. A., Bell, R. M., and Kinkade, J. M., Jr. (1986) Inhibition of phorbol ester-dependent differentiation of human promyelocytic leukemic (HL-60) cells by sphinganine and other long-chain bases. J Biol Chem 261,12610-5 Merrill, A. H., Jr., and Wang, E. (1992) Enzymes of ceramide biosynthesis. Methods Enzymal 209,427-3 7. Merrill, A. H., Jr., Wang, E., Vales, T. R., Smith, E. R., Schroeder, J. J ., Menaldino, D. 8., Alexander, C., Crane, H. M., Xia, J ., Liotta, D. C., Meredith, F. 1., and Riley, R. T. (1996) F umonisin toxicity and sphingolipid biosynthesis. Adv Exp Med Biol 392,297-306 Micoli, K. J., Pan, G., Wu, Y., Williams, J. P., Cook, W. J., and McDonald, J. M. (2000) Requirement of calmodulin binding by HIV-1 gp160 for enhanced F AS-mediated apoptosis. J Biol Chem 275,123 3-40 128 Miller-Podraza, H., Bradley, R. M., and F ishman, P. H. (1982) Biosynthesis and localization of gangliosides in cultured cells. Biochemistry 21,3260-5. Morgan, M. K., Schroeder, J. J., Rottinghaus, G. E., Powell, D. C., Bursian, S. J., and Aulerich, R. J. (1997) Dietary fumonisins disrupt sphingolipid metabolism in mink and increase the free sphinganine to sphingosine ratio in urine but not in hair. Vet Hum Toxicol 39,334-6 Motelin, G. K., Haschek, W. M., Ness, D. K., Hall, W. F., Harlin, K. S., Schaeffer, D. J ., and Beasley, V. R. (1994) Temporal and dose-response features in swine fed corn screenings contaminated with fumonisin mycotoxins. Mycopathologia 126,27-40 Nikolova—Karakashian, M., Morgan, E. T., Alexander, C., Liotta, D. C., and Merrill, A. H., Jr. (1997) Bimodal regulation of ceramidase by interleukin-1 beta. Implications for the regulation of cytochrome p450 2C1 1. J Biol Chem 272,18718-24. Nilsson, A. (1969) The presence of spingomyelin- and ceramide-cleaving enzymes in the small intestinal tract. Biochim Biophys Acta 176,339-47. Nishida, E., and Gotoh, Y. (1993) The MAP kinase cascade is essential for diverse signal transduction pathways. Trends Biochem Sci 18,128-31. Norred, W. P., Plattner, R. D., Vesonder, R. F., Bacon, C. W., and Voss, K. A. (1992) Effects of selected secondary metabolites of Fusarium moniliforme on unscheduled synthesis of DNA by rat primary hepatocytes. Food Chem Toxicol 30,233-7 Norred, W. P., Plattner, R. D., and Chamberlain, W. J. (1993) Distribution and excretion of [14C] fumonisin B1 in male Sprague-Dawley rats. Nat toxins. New York, NY: Wiley Liss, Inc 341-346 Norred, W. P., Plattner, R. D., Vesonder, R. F., Bacon, C. W., and Voss, K. A. (1992) Effects of selected secondary metabolites of Fusarium moniliforme on unscheduled synthesis of DNA by rat primary hepatocytes. F aod Chem Toxicol 30,233-7. Obeid, L. M., Linardic, C. M., Karolak, L. A., and Hannun, Y. A. (1993) Programmed cell death induced by ceramide. Science 259,1769-71 Ohnishi, M., and Fujino, Y. (1982) Sphingolipids in immature and mature soybeans. Lipids 17, 803-810 Ohta, H., Sweeney, E. A., Masamune, A., Yatomi, Y., Hakomori, S., and Igarashi, Y. (1995) Induction of apoptosis by sphingosine in human leukemic HL-60 cells: a possible endogenous modulator of apoptotic DNA fragmentation occurring during phorbol ester- induced differentiation. Cancer Res 55, 691-7 Ohta, H., Yatomi, Y., Sweeney, E. A., Hakomori, S., and Igarashi, Y. (1994) A possible role 129 of sphingosine in induction of apoptosis by tumor necrosis factor-alpha in human neutrophils. F EBS Lett 355,267-70. Oishi, K., Raynor, R. L., Charp, P. A., and Kuo, J. F. (1988) Regulation of protein kinase C by lysophospholipids. Potential role in signal transduction. J Biol Chem 263,6865-71. Okazaki, T., Bielawska, A., Bell, R. M., and Hannun, Y. A. (1990) Role of ceramide as a lipid mediator of 1 alpha,25-dihydroxyvitamin D3-induced HL-60 cell differentiation. J Biol Chem 265,15823-31 Okazaki, T., Bell, R. M., and Hannun, Y. A. (1989) Sphingomyelin turnover induced by vitamin D3 in HL-60 cells. Role in cell differentiation. J Biol Chem 264,19076-80 Olivera, A., Edsall, L., Poulton, S., Kazlauskas, A., and Spiegel, S. (1999a) Platelet-derived growth factor-induced activation of sphingosine kinase requires phosphorylation of the PDGF receptor tyrosine residue responsible for binding of PLCgamma. FASEB J 13,1593- 600 Olivera, A., Kohama, T., Edsall, L., Nava, V., Cuvillier, O., Poulton, S., and Spiegel, S. (1999b) Sphingosine kinase expression increases intracellular sphingosine- 1 -phosphate and promotes cell growth and survival. J Cell Biol 147,545-58 Olivera, A., Zhang, H., Carlson, R. O., Mattie, M. E., Schmidt, R. R., and Spiegel, S. (1994) Stereospecificity of sphingosine-induced intracellular calcium mobilization and cellular proliferation. J Biol Chem 269,17924-30 Olivera, A., and Spiegel, S. (1993) Sphingosine-l-phosphate as second messenger in cell proliferation induced by PDGF and FCS mitogens. Nature 365,557-60 Osawa M, Swindells MB, Tanikawa 1, Tanaka T, Mase T, Furuya T, Ikura M (1998) Solution structure of calmodulin-W-7 complex: the basis of diversity in molecular recognition. J. Mol. Biol. 276, 165-76 Osweiler, G. D., Ross, P. F., Wilson, T. M., Nelson, P. E., Witte, S. T., Carson, T. L., Rice, L. G., and Nelson, H. A. (1992) Characterization of an epizootic of pulmonary edema in swine associated with fumonisin in corn screenings. J Vet Diagn Invest 4,53-9. Pan, G., Zhou, T., Radding, W., Saag, M. S., Mountz, J. D., and McDonald, J. M. (1998) Calmodulin antagonists inhibit apoptosis of CD4+ T-cells from patients with AIDS. Immunopharmacology 40,91 -1 03. Pan, X., Solomon, S. S., Shah, R. J ., Palazzolo, M. R., and Raghow, R. S. (2000) Members of the Sp transcription factor family regulate rat calmodulin gene expression. J Lab C [in Med 136,157-63. 130 i‘eL-um .fi 1‘. A I 1‘ Pelech, S. L., Charest, D. L., Mordret, G. P., Siow, Y. L., Palaty, C., Campbell, D., Charlton, L., Samiei, M., and Sanghera, J. S. (1993) Networking with mitogen-activated protein kinases. Mol Cell Biochem 127-128,157-69. Pinelli, E., Poux, N., Garren, L., Pipy, B., Castegnaro, M., Miller, D. J., and Pfohl- Leszkowicz, A. (1999) Activation of mitogen-activated protein kinase by fumonisin B, stimulates cPLA2 phosphorylation, the arachidonic acid cascade and cAMP production. Carcinogenesis 20,1683-8. Pinton, P., Ferrari, D., Rapizzi, E., Di Virgilio, F. D., Pozzan, T., and Rizzuto, R. (2001) The Ca 2* concentration of the endoplasmic reticulum is a key determinant of ceramide-induced apoptosis: significance for the molecular mechanism of Bcl-2 action. Embo J 20,2690-701. Plattner, R. D., Ross, P. F ., Reagor, J ., Stedelin, J ., and Rice, L. G. (1991) Analysis of corn and cultured corn for fumonisin B, by HPLC and GC/MS by four laboratories. J Vet Diagn Invest 3,357-8 Prelusky, D. B., Miller, J. D., and Trenholm, H. L. (1996) Disposition of 14C-derived residues in tissues of pigs fed radiolabeled fumonisin B,. Food Addit Contam 13,155-62 Prelusky, D. B,, Trenholm, H. L., and Savard, M. E. (1994) Pharmacokinetic fate of 14C- labelled fumonisin B, in swine. Nat T axins 2,73-80 Preston Martin, 8., Pike, M. C., Ross, R. K., Jones, P. A., and Henderson, B. E. (1990) Increased cell division as a cause of human cancer. Cancer Res 50,7415-21 Pushkareva, M., Chao, R., Bielawska, A., Merrill, A. H., Crane, H. M., Lagu, B., Liotta, D., and Hannun, Y. A. (1995) Stereoselectivity of induction of the retinoblastoma gene product (pr) dephosphorylation by D-erythro-sphingosine supports a role for pr in growth suppression by sphingosine. Biochemistry 34,1885-92. Putzer, B. M., Stiewe, T., Parssanedjad, K., Rega, S., and Esche, H. (2000)E1A is sufficient by itself to induce apoptosis independent of p53 and other adenoviral gene products. Cell Death Differ 7,177-88 Raeder, E. M., Mansfield, P. J., Hinkovska-Galcheva, V., Kjeldsen, L., Shayman, J. A., and Boxer, L. A. (1999) Sphingosine blocks human polymorphonuclear leukocyte phagocytosis through inhibition of mitogen-activated protein kinase activation. Blood 93,686-93. Raines, M. A., Kolesnick, R. N., and Golde, D. W. (1993) Sphingomyelinase and ceramide activate mitogen-activated protein kinase in myeloid HL-60 cells. J Biol Chem 268,145 72-5 Ramljak, D., Calvert, R. J ., Wiesenfeld, P. W., Diwan, B. A., Catipovic, B., Marasas, W. F ., Victor, T. C., Anderson, L. M., and Gelderblom, W. C. (2000) A potential mechanism for fumonisin Bl-mediated hepatocarcinogenesis: cyclin D1 stabilization associated with activation of Akt and inhibition of GSK-3beta activity. Carcinogenesis 21,1537-46 131 Rasmussen, C. D., Lu, K. P., Means, R. L., and Means, A. R. (1992) Calmodulin and cell cycle control. J Physiol Paris 86,83-8. Rasmussen, C. D., and Means, A. R. (1989) Calmodulin, cell growth and gene expression. Trends Neurosci 12,433-8. Rasmussen, C. D., and Means, A. R. (1987) Calmodulin is involved in regulation of cell proliferation. Embo J 6,3961 -8. Restum, J. C., Bursian, S. J., Millerick, M., Render, J. A., Merrill, A. H., Wang, E., Rottinghaus, G. E., and Aulerich, R. J. (1995) Chronic toxicity of fumonisins from F usarium moniliforme culture material (M-l325) to mink. Arch Environ Contam T oxicol 29,545-50. Riley, R. T., Hinton, D. M., Chamberlain, W. J ., Bacon, C. W., Wang, E., Merrill, A. H., Jr., and Voss, K. A. (1994) Dietary fumonisin B, induces disruption of sphingolipid metabolism in Sprague-Dawley rats: a new mechanism of nephrotoxicity. J Nutr 124,594-603 Riley, R. T., An, N. H., Showker, J. L., Yoo, H. S., Norred, W. P., Chamberlain, W. J., Wang, E., Menill, A. H., Motelin, G., Beasley, V. R., and et al. (1993a) Alteration of tissue and serum sphinganine to sphingosine ratio: an early biomarker of exposure to fumonisin- containing feeds in pigs. T axical Appl Pharmacol 118,105-12. Riley, R. T., Norred, W. P., and Bacon, C. W. (1993b) Fungal toxins in foods: recent concerns. Annu Rev Nutr 13,167-89 Riley, R. T., Voss, K. A., Norred, W. P., Bacon, C. W., Meredith, F. 1., and Shanna, RP. (1999) Serine palmitoyltransferase inhibition reverses anti-proliferative effects of ceramide synthase inhibition in cultured renal cells and suppresses free sphingoid base accumulation in kidney of BALBc mice. Environ T axical Pharmacol 7,109-118 Riley, R. T., Voss, K. A., Yoo, H. S., Gelderblom, W. C. A., and Merrill, A. H., Jr. (1994b) Mechanism of fumonisin toxicity and carcinogenesis. J food prot. Des Moines, Iowa .' International Association of Milk, Food and Environmental Sanitarians. July 63 8-645 Rose, E. F. (1981) A review of factors associated with cancer of the esophagus in Transkei. Prog Clin Biol Res 53,67-75. Rosenthal, D. S., Simbulan Rosenthal, C. M., Iyer, S., Spoonde, A., Smith, W., Ray, R., and Smulson, M. E. (1998) Sulfur mustard induces markers of terminal differentiation and apoptosis in keratinocytes via a Cay-calmodulin and caspase-dependent pathway. J Invest Dermatol 1 1 1 ,64-7 1 Ross, P. F ., Rice, L. G., Osweiler, G. D., Nelson, P. E., Richard, J. L., and Wilson, T. M. (1992) A review and update of animal toxicoses associated with fumonisin-contaminated feeds and production of fumonisins by Fusarium isolates. Mycopathologia 117,109-14 132 Ross, P. F., Rice, L. G., Plattner, R. D., Osweiler, G. D., Wilson, T. M., Owens, D. L., Nelson, H. A., and Richard, J. L. (1991) Concentrations of fumonisin B, in feeds associated with animal health problems. Mycopathologia 114,129-35 Rother, J., van Echten, G., Schwarzrnann, G., and Sandhoff, K. (1992) Biosynthesis of sphingolipids: dihydroceramide and not sphinganine is desaturated by cultured cells. Biochem Biophys Res Commun 189,14-20. Saba, J. D., Obeid, L. M., and Hannun, Y. A. (1996) Ceramide: an intracellular mediator of apoptosis and growth suppression. Philos Trans R Soc Land B Biol Sci 351,233-40 Saito, M., Terui, Y., and Nojiri, H. (1985) An acidic glycosphingolipid, monosialo- ganglioside GM3, is a potent physiological inducer for monocytic differentiation of human promyelocytic leukemia cell line HL-60 cells. Biochem Biophys Res Commun 132,223-31. Sakakura, C., Sweeney, E. A., Shirahama, T., Hakomori, S., and Igarashi, Y. (1996) Suppression of bcl-2 gene expression by sphingosine in the apoptosis of human leukemic HL-60 cells during phorbol ester-induced terminal differentiation. F EBS Lett 379,177-80 Sakakura, C., Sweeney, E. A., Shirahama, T., Hagiwara, A., Yamaguchi, T., Takahashi, T., Hakomori, S., and Igarashi, Y. (1998) Selectivity of sphingosine-induced apoptosis. Lack of activity of DL-erythyro-dihydrosphingosine. Biochem Biophys Res C ommun 246,827-30 Schmelz, E. M., Crall, K. J., Larocque, R., Dillehay, D. L., and Merrill, A. H., Jr. (1994) Uptake and metabolism of sphingolipids in isolated intestinal loops of mice. J Nutr 124,702- 12. Schrarnek, H., Feifel, E., Healy, E., and Pollack, V. (1997) Constitutively active mutant of the mitogen-activated protein kinase kinase MEKl induces epithelial dedifferentiation and growth inhibition in madin-darby canine kidney-C7 cells. J Biol Chem 272,11426-33. Schroeder, J. J., Crane, H. M., Xia, J., Liotta, D. C., and Merrill, A. H. (1994) Disruption of sphingolipid metabolism and stimulation of DNA synthesis by fumonisin B,. A molecular mechanism for carcinogenesis associated with Fusarium moniliforme. J Biol Chem 269,3475-81. Schroeder, J. J., and Cousins, R. J. (1990) Interleukin 6 regulates metallothionein gene expression and zinc metabolism in hepatocyte monolayer cultures. Proc Natl Acad Sci U S A 87,3137-41 Schubert, K. M., Scheid, M. P., and Duronio, V. (2000) Ceramide inhibits protein kinase B/Akt by promoting dephosphorylation of serine 473. J Biol Chem 275,13330-5. Schwarz, A., Rapaport, E., Hirschberg, K., and F uterrnan, A. H. (1995) A regulatory role for sphingolipids in neuronal growth. Inhibition of sphingolipid synthesis and degradation have opposite effects on axonal branching. J Biol Chem 270,10990-8 133 Sellins, KS, and Cohen, J .J . (1995) Nuclear changes in the cytotoxic T lymphocyte-induced model of apoptosis. Immunol. Rev. 146, 241-66 Shelby, R. A., White, D. G., and Bauske, E. M. (1994) Differential fumonisin production in maize hybrids. Plant dis. [St Paul, Minn, American Phytopathological Society]. June 582- 584 Shephard, G. S., Thiel, P. G., Sydenham, E. W., and Savard, M. E. (1995) Fate of a single dose of l‘lC-labelled fumonisin B, in vervet monkeys. Nat Toxins 3,145-50 Shephard, G. S., Van Der Westhuizen, L., Thiel, P. G., Gelderblom, W. C. A., Marasas, W. F. O., and Van Schalkwyk, D. J. (1996) Disruption of sphingolipid metabolism in non- human primates consuming diets of fumonisin-containing Fusarium moniliforme culture material. Toxicon. Oxford: Elsevier Science Ltd. May 527-534 Shephard, G. S., Thiel, P. G., Sydenham, E. W., Vleggaar, R., and Alberts, J. F. (1994c) Determination of the mycotoxin fumonisin B, and identification of its partially hydrolysed metabolites in the faeces of non-human primates. Food Chem Toxicol 32,23-9 Shephard, G. S., Thiel, P. G., Sydenham, E. W., Alberts, J. F., and Cawood, M. E. (1994b) Distribution and excretion of a single dose of the mycotoxin fumonisin B, in a non-human primate. T axican 32,73 5-41 Shephard, G. S., Thiel, P. G., Sydenham, E. W., and Alberts, J. F. (1994a) Biliary excretion of the mycotoxin fumonisin B, in rats. F oad Chem Toxicol 32,489-91 Shirahama, T., Sakakura, C., Sweeney, E. A., Ozawa, M., Takemoto, M., Nishiyama, K., Ohi, Y., and Igarashi, Y. (1997) Sphingosine induces apoptosis in androgen-independent human prostatic carcinoma DU-l45 cells by suppression of bcl-X, gene expression. FEBS Lett 407,97-100 Singer, A. L., Sherwin, R. P., Dunn, A. S., and Appleman, M. M. (1976) Cyclic nucleotide phosphodiesterases in neoplastic and nonneoplastic human mammary tissues. Cancer Res 36,60-6. Smith, E. R., Menill, A. H., Obeid, L. M., and Hannun, Y. A. (2000) Effects of sphingosine and other sphingolipids on protein kinase C. Methods Enzymol 312,361-73. Smith, E. R., and Merrill, A. H., Jr. (1995) Differential roles of de nova sphingolipid biosynthesis and turnover in the "burst" of free sphingosine and sphinganine, and their 1- phosphates and N-acyl-derivatives, that occurs upon changing the medium of cells in culture. J Biol Chem 270,18749-58 Smyth, M. J., Obeid, L. M., and Hannun, Y. A. (1997) Ceramide: a novel lipid mediator of apoptosis. Adv Pharmacol 41,133-54 134 Solomon, S. S., Palazzolo, M. R., Takahashi, T., and Raghow, R. (1997) Transcription factor Spl is necessary for basal calmodulin gene transcription and for its selective stimulation by insulin. Endocrinology 138,5052-4. Spiegel, S., and Merrill, A. H., Jr. (1996) Sphingolipid metabolism and cell growth regulation. FASEB J 10,1388-97 Spiegel, S., Olivera, A., and Carlson, R. O. (1993) The role of sphingosine in cell growth regulation and transmembrane signaling. Adv Lipid Res 25,105-29 Spiegel, S., and F ishman, P. H. (1987) Gangliosides as bimodal regulators of cell growth. Prac Natl Acad Sci USA 84,141-5. Spiegel, S., Fishman, P. H., and Weber, R. J. (1985) Direct evidence that endogenous GMl ganglioside can mediate thymocyte proliferation. Science 230,1285-7. Spiegel, S. (1999) Sphingosine l-phosphate: a prototype of a new class of second messengers. J Leukoc Biol 65,341-4 Spiegel, S., Olivera, A., Zhang, H., Thompson, E. W., Su, Y., and Berger, A. (1994) Sphingosine-1 -phosphate, 3 novel second messenger involved in cell growth regulation and signal transduction, affects growth and invasiveness of human breast cancer cells. Breast Cancer Res Treat 31,337-48 Stemmer, P. M., and Klee, C. B. (1994) Dual calcium ion regulation of calcineurin by calmodulin and calcineurin B. Biochemistry 33,6859-66. Stephens, L., Anderson, K., Stokoe, D., Erdjument Bromage, H., Painter, G. F., Holmes, A. B., Gaffney, P. R., Reese, C. B., McCormick, F ., Tempst, P., Coadwell, J ., and Hawkins, P. T. (1998) Protein kinase B kinases that mediate phosphatidylinositol 3,4,5-trisphosphate- dependent activation of protein kinase B. Science 279,710-4. Stevens, V. L., Nimkar, S., Jamison, W. C., Liotta, D. C., and Merrill, A. H., Jr. (1990) Characteristics of the growth inhibition and cytotoxicity of long-chain (sphingoid) bases for Chinese hamster ovary cells: evidence for an involvement of protein kinase C. Biochim Biophys Acta 1051,3 7-45. Stoclet, J. C., Gerard, D., Kilhoffer, M. C., Lugnier, C., Miller, R., and Schaeffer, P. (1987) Calmodulin and its role in intracellular calcium regulation. Prog Neurobiol 29,321-64. Stoffel, W., Sticht, G., and LeKim, D. (1968) Metabolism of sphingosine bases. IX. Degradation in vitro of dihydrospingosine and dihydrospingosine phosphate to palmitaldehyde and ethanolamine phosphate. Hoppe Seylers Z Physiol Chem 349,1745-8. Stoffel, W., and Assmann, G. (1970) Metabolism of sphingosine bases. XV. Enzymatic degradation of 4t-sphingenine l-phosphate (sphingosine l-phosphate) to 2t-hexadecen-1 -al 135 and ethanolamine phosphate. Hoppe Seylers Z Physiol Chem 351,1041-9. Stokoe, D., Stephens, L. R., Copeland, T., Gaffney, P. R., Reese, C. B., Painter, G. F ., Holmes, A. B., McCormick, F., and Hawkins, P. T. (1997) Dual role of phosphatidylinositol- 3,4,5-trisphosphate in the activation of protein kinase B. Science 277,567-70. Su, Y., Rosenthal, D., Smulson, M., and Spiegel, S. (1994) Sphingosine l-phosphate, a novel signaling molecule, stimulates DNA binding activity of AP-l in quiescent Swiss 3T3 fibroblasts. J Biol Chem 269,16512-7 Sydenham, E. W., Gelderblom, W. C. A., Thiel, P. G., and Marasas, W. F. O. (1990) Evidence for the natural occurrence of fumonisin B,, a mycotoxin produced by Fusarium moniliforme, in corn. J A gric Food Chem. Washington, D. C. .' American Chemical Society. Jan 285-290 Telford, W. G., King, L. E., and F raker, P. J. (1994) Rapid quantitation of apoptosis in pure and heterogeneous cell populations using flow cytometry. J Immunol Methods 172,1 -16 Tolleson, W. H., Dooley, K. L., Sheldon, W. G., Thurman, J. D., Bucci, T. J., and Howard, P. C. (1996a) The mycotoxin fumonisin induces apoptosis in cultured human cells and in livers and kidneys of rats. Adv Exp Med Biol 392,23 7-50 Tolleson, W. H., Melchior, W. B., Morris, S. M., McGarrity, L. J., Domon, O. E., Muskhelishvili, L., James, S. J ., and Howard, P. C. (1996b) Apoptotic and anti-proliferative effects of fumonisin B, in human keratinocytes, fibroblasts, esophageal epithelial cells and hepatoma cells. Carcinogenesis 17,239-49. Tolleson, W. H., Couch, L. H., Melchior, W. B., Jenkins, G. R., Muskhelishvili, M., Muskhelishvili, L., McGarrity, L. J., Domon, 0., Morris, S. M., and Howard, P. C. (1999) Fumonisin B, induces apoptosis in cultured human keratinocytes through sphinganine accumulation and ceramide depletion. Int J Oncol 14,833-43. Tomquist, K., Saarinen, P., Vainio, M., and Ahlstrom, M. (1997) Sphingosine l-phosphate mobilizes sequestered calcium, activates calcium entry, and stimulates deoxyribonucleic acid synthesis in thyroid F RTL-5 cells. Endocrinology 138,4049-57 Toutenhoofd, S. L., Foletti, D., Wicki, R., Rhyner, J. A., Garcia, F., Tolon, R., and Strehler, E. E. (1998) Characterization of the human CALM2 calmodulin gene and comparison of the transcriptional activity of CALM], CALM2 and CALM3. Cell Calcium 23,323-38. Tsatsanis, C., and Spandidos, D. A. (2000) The role of oncogenic kinases in human cancer. Int J Mol Med 5,5 83-90. Tuccari, G., Rizzo, A., and Barresi, G. (1993) Immunocytochemical demonstration of calmodulin and its activated forms in normal and carcinomatous human mammary tissue. Eur J Histachem 37,59-64. 136 Ueno, Y., Iijima, K., Wang, S. D., Sugiura, Y., Sekijima, M., Tanaka, T., Chen, C., and Yu, S. Z. (1997) F umonisins as a possible contributory risk factor for primary liver cancer: a 3- year study of corn harvested in Haimen, China, by HPLC and ELISA. Food Chem Toxicol 35,1143-50. Van Der Westhuizen, L., Shephard, G. S., Snyman, S. D., Abel, S., Swanevelder, S., and Gelderblom, W. C. A. (1998) Inhibition of sphingolipid biosynthesis in rat primary hepatocyte cultures by fumonisin B, and other structurally related compounds. Food chem toxicol. Oxford, UK. .' Elsevier Science Ltd. June 497-503 Van Eldik, L. J ., and Burgess, W. H. (1983) Analytical subcellular distribution of calmodulin and calmodulin-binding proteins in normal and virus-transforrned fibroblasts. J Biol Chem 258,4539-47. Venable, M. E., Blobe, G. C., and Obeid, L. M. (1994) Identification of a defect in the phospholipase D/diacylglycerol pathway in cellular senescence. J Biol Chem 269,26040-4. Veigl, M. L., Sedwick, W. D., and Vanaman, T. C. (1982) Calmodulin and Ca 2* in normal and transformed cells. Fed Proc 41,2283-8. Voss, K. A., Riley, R. T., Norred, W. P., Bacon, C. W., Meredith, F. 1., Howard, P. C., Plattner, R. D., Collins, T. F., Hansen, D. K., and Porter, J. K. (2001) An overview of rodent toxicities: liver and kidney effects of fumonisins and F usarium moniliforme. Environ Health Perspect 109,259-66. Voss, K. A., Chamberlain, W. J., Bacon, C. W., and Norred, W. P. (1993) A preliminary investigation on renal and hepatic toxicity in rats fed purified fumonisin B,. Nat Toxins 1,222-8 Voss, K. A., Riley, R. T., Norred, W. P., Bacon, C. W., Meredith, F. 1., Howard, P. C., Plattner, R. D., Collins, T. F ., Hansen, D. K., and Porter, J. K. (2001) An overview of rodent toxicities: liver and kidney effects of firmonisins and F usarium moniliforme. Environ Health Perspect 109,259-66. Vudathala, D. K., Prelusky, D. B., Ayroud, M., Trenholm, H. L., and Miller, J. D. (1994) Pharrnacokinetic fate and pathological effects of l“C-fumonisin B, in laying hens. Nat Toxins 2,81-8 Wang, H. G., Pathan, N., Ethell, I. M., Krajewski, S., Yamaguchi, Y., Shibasaki, F., McKeon, F., Bobo, T., F ranke, T. F ., and Reed, J. C. (1999) Ca2+-induced apoptosis through calcineurin dephosphorylation of BAD. Science 284,339-43 Wang, W., Jones, C., Ciacci-Zanella, J., Holt, T., Gilchrist, D. G., and Dickman, M. B. (1996) F umonisins and Alternaria alternata lycapersici toxins: sphinganine analog mycotoxins induce apoptosis in monkey kidney cells. Proc Natl Acad Sci U S A 93,3461-5. 137 Wang, D. G., Johnston, C. F., Marley, J. J., Phenix, K. V., Atkinson, A. 8., Russell, C. F ., and Buchanan, K. D. (1997) Expression of the apoptosis-suppressing gene BCL-2 in pheochromocytoma is associated with the expression of C-MYC. J Clin Endocrinol Metab 82,1949-52 Wang, E., Ross, P. F., Wilson, T. M., Riley, R. T., and Merrill, A. H., Jr. (1992) Increases in serum sphingosine and sphinganine and decreases in complex sphingolipids in ponies given feed containing fumonisins, mycotoxins produced by Fusarium moniliforme. J Nutr 122,1706-16. Wang, T. T., and Phang, J. M. (1995) Effects of estrogen on apoptotic pathways in human breast cancer cell line MCF-7. Cancer Res 55,2487-9 Wang, E., Norred, W. P., Bacon, C. W., Riley, R. T., and Merrill, A. H., Jr. (1991) Inhibition of sphingolipid biosynthesis by fumonisins. Implications for diseases associated with Fusarium moniliforme. J Biol Chem. 266,14486-14490 Wattenberg, E. V., Badria, F. A., and Shier, W. T. (1996) Activation of mitogen-activated protein kinase by the carcinogenic mycotoxin fumonisin B,. Biochem Biophys Res Commun 227,622-7. Wei, J. W., Morris, H. P., and Hickie, R. A. (1982) Positive correlation between calmodulin content and hepatoma growth rates. Cancer Res 42,2571-4. Wieder, T., Geilen, C. C., Kolter, T., Sadeghlar, F ., Sandhoff, K., Brossmer, R., Ihrig, P., Perry, D., Orfanos, C. E., and Hannun, Y. A. (1997) Bel-2 antagonizes apoptotic cell death induced by two new ceramide analogues. FEBS Lett 411,260-4 Wilson, E., Olcott, M. C., Bell, R. M., Merrill, A. H., Jr., and Lambeth, J. D. (1986) Inhibition of the oxidative burst in human neutrophils by sphingoid long-chain bases. Role of protein kinase C in activation of the burst. J Biol Chem 261,12616-23. Wilson, T. M., Ross, P. F., Rice, L. G., Osweiler, G. D., Nelson, H. A., Owens, D. L., Plattner, R. D., Reggiardo, C., Noon, T. H., and Pickrell, J. W. (1990) Fumonisin B, levels associated with an epizootic of equine leukoencephalomalacia. J Vet Diagn Invest 2,213-6 Wu, J ., Spiegel, S., and Sturgill, T. W. (1995) Sphingosine l-phosphate rapidly activates the mitogen-activated protein kinase pathway by a G protein-dependent mechanism. J Biol Chem 270,1 1484-8 Yao, B., Zhang, Y., Delikat, S., Mathias, S., Basu, S., and Kolesnick, R. (1995) Phosphorylation of Raf by ceramide-activated protein kinase. Nature 378,307-10 Yoo, H. S., Norred, W. P., Showker, J ., and Riley, R. T. (1996) Elevated sphingoid bases and complex sphingolipid depletion as contributing factors in fumonisin-induced cytotoxicity. Toxicol Appl Pharmacol 138,21 1-8 138 Yoo, H. S., Norred, W. P., Wang, E., Merrill, A. H., Jr., and Riley, R. T. (1992) Fumonisin inhibition of de novo sphingolipid biosynthesis and cytotoxicity are correlated in LLC-PK, cells. Toxicol Appl Pharmacol. 114, 9-15 Yuan, Z. Q., Sun, M., F eldman, R. 1., Wang, G., Ma, X., Jiang, C., Coppola, D., Nicosia, S. V., and Cheng, J. Q. (2000) Frequent activation of AKT2 and induction of apoptosis by inhibition of phosphoinositide-B-OH kinase/Akt pathway in human ovarian cancer. Oncogene 19,2324-30. Zeisel, S. H., Char, D., and Sheard, N. F. (1986) Choline, phosphatidylcholine and sphingomyelin in human and bovine milk and infant formulas. J Nutr 116,50-8. Zhang, H., Buckley, N. E., Gibson, K., and Spiegel, S. (1990) Sphingosine stimulates cellular proliferation via a protein kinase C-independent pathway. J Biol Chem 265,76-81 Zhang, H., Desai, N. N., Olivera, A., Seki, T., Brooker, G., and Spiegel, S. (1991) Sphingosine-l-phosphate, a novel lipid, involved in cellular proliferation. J Cell Biol 114,155-67 Zhang, Y., Dickman, M. B., and Jones, C. (1999) The mycotoxin fumonisin B1 transcriptionally activates the p21 promoter through a cis-acting element containing two Spl binding sites. J Biol Chem 274,12367-71 Zhang, J ., Alter, N., Reed, J. C., Bomer, C., Obeid, L. M., and Hannun, Y. A. (1996) Bel-2 interrupts the ceramide-mediated pathway of cell death. Prac Natl Acad Sci U S A 93,5325-8 Zhou, H., Li, X. M., Meinkoth, J., and Pittman, R. N. (2000) Akt regulates cell survival and apoptosis at a postrnitochondrial level. J Cell Biol 151,483-94. Zundel, W., and Giaccia, A. (1998) Inhibition of the anti-apoptotic PI3K/Akt/Bad pathway by stress. Genes Dev 12,1941-6. 139